You are on page 1of 23

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Copyright 1997 Elsevier Science Ltd Int. J. Rock Mech. & Min. Sci. Vol. 34, No. 3-4, 1997 ISSN 0148-9062 To cite this paper: Int. J. RockMech. &Min. Sci. 34:3-4, Paper No. 112

C O M P R E S S I V E FAILURE OF R O C K S
V i j a y G u p t a l ; J 6 r g e n S. B e r g s t r 6 m 2

1 Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, CA 90095, USA z Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
ABSTRACT

A model for the compressive failure of rocks via the process of shear faulting is presented. The model addresses the progressive growth of damage that leads to the formation of a critical fault nucleus, which grows unstably in its own plane by fracturing the grain boundaries in an increasingly rapid succession. The model uses a two parameter Weibull-type shear strength distribution for the defining nucleation of initial damage, followed by the use of stress enhancement factors for addressing the increased probability of failure in the vicinity of already cracked grain boundaries. As the stress is further increased, similar correlated fracturing events gets preferentially aligned to the crack cluster, resulting in an echelon of cracks. This crack cluster is modeled as an elliptical inhomogeneity with a lower shear modulus compared with the uncracked material on the outside. The shear stress concentration resulting from this moduli mismatch was calculated and used to compute the stress enhancement factors for defining the nucleation of additional cracking events near the crack cluster. Eventually, the size of the crack cluster becomes sufficiently large such that it carries a stress concentration high enough to fracture all grain boundary elements in front of it in an increasingly rapid succession. The stress associated with this event is taken as the failure stress. The model is applied to a variety of rocks, including granite, eclogite, gabbro, aplite, rocksalt, sandstone, dunite, limestone, and marble, and the results compare rather well with the failure data from the literature.
Copyright 1997 Elsevier Science Ltd

KEYWORDS Compressive strength Constitutive Relations Shear Faults Brittle Failure Triaxial Testing Statistical Analysis Shear Strength Loading Mechanical Properties Drilling Slope Stability INTRODUCTION

It is now widely recognized (Horii, Nemat-Nasser 1985,1986; Ashby, Hallam 1986; Griggs, Handin 1960; Fairhurst, Cook 1966; Paterson 1978; and Kranz 1983) that under critical conditions of temperature, strain rate, and lateral confining pressure, brittle solids like rock fail via one of the following modes: (1) axial splitting of the sample in the direction of principal compression at low or non-existing confining pressure; (2) faulting or macroscopic shear failure under moderate confining pressures; and finally, (3) ductile flow in the presence of a suitably large confining pressure. The micromechanical events leading to the above failure modes are somewhat similar. The failure starts

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

by nucleation of several transgranular and grain boundary cracks almost quasi homogeneously within the rock volume. The transgranular cracks, which are aligned nearly parallel with the maximum compressive stress (Tapponier, Brace 1976) and typically traverse an entire grain, are observed to nucleate from pores within the grains, from high angle grain boundaries of different minerals, and from pre-existing grain boundary cavities. A large number of these microcracks are associated with point contacts between grains and with grain boundaries between minerals of greatly different elastic moduli (Tapponier, Brace 1976; Kranz 1983; Batzle et al. 1979). The grain boundary cracks on the other hand initiate from pre-existing arrays of low aspect ratio boundary cavities, healed fissures and second phase particles. At low confinement, the transgranular cracks grow long and then coalesce with their neighbors causing macroscopic failure in the form of splitting. Application of moderate confining pressures severely limits crack extension and thus inhibits axial splitting. This allows the sample to sustain additional stress which results in new cracking events, many of which occur near previously nucleated cracks. Additional loading results in new cracks to get preferentially aligned with existing cracks and soon the local damage takes the form of a crack cluster. During further loading, besides additional cracking, the damage in previously cracked grain intensifies on account of transgranular cracks leading to grain pulverization, rotation and compaction. All these events acts as stimuli for concentrating stress and enhances the probability of preferentially cracking. At a critical load, one of these crack clusters becomes unstable and takes the form of a shear fault. As expected, a section through the sample reveals coalescence of arrays and networks of grain boundary cracks, transgranular cracks and pores along the fault plane (Wong 1982). The progressive failure process discussed above was directly established by monitoring acoustic emissions from new cracking events in instrumented rock samples (Schulz 1968; Bailey et al. 1979; Rothman 1977). Experiments further show that several of such clusters can be formed, but only one of them grows to the final failure suggesting the influence of the statistical strength variability within the sample. Under high confining pressures, the growth of wing cracks is dramatically suppressed and the material shows enhanced resistance to frictional sliding. This leads to a delay in the formation of the critical fault nucleus, during which, the rock undergoes isolated fractures quite homogeneously over its entire volume. This homogeneous cracking is essentially due to the formation of several simultaneous sub-critical fault nuclei. In the absence of failure via a terminal shear fault, the high confinement sample fails by undergoing large macroscopic dilatation similar to those observed in ductile solids. The above failure process is largely observed in rocks of different types, although the exact stress and confinement ratios for failure mode transition can vary significantly. Similarly the details of the crack density and the resulting coalescence process can also change. The failure in all the three regimes has been modeled using both micromechanical and continuum approaches by a number of investigators (Horii, Nemat-Nasser 1985,1986; Ashby, Hallam 1986; Kachanov 1982; Nemat-Nasser, Obata 1988, and Rudnicki, Rice 1975). Due to space limitation, a detailed review of these models is deferred to a more expanded version of this article (Gupta, Bergstr6m 1997) Here we only note that unlike previous approaches, the present work addresses the nucleation and progressive growth of the critical fault nucleus that spreads unstably to cause the failure of the rock sample. The progressive growth of the fault nucleus is considered in a statistical manner by combining the stress concentration at the fault nucleus with an assumed distribution of weak grain boundaries within the rock volume. The model is exemplified by applying it to a wide variety of rocks, including granite, eclogite, gabbro, aplite, rocksalt, sandstone, dunite, limestone, and marble. Although we focus on failure by shear faulting, for the sake of completion and for comparing failure data obtained under largely varying confinements, we present the model of Ashby, Hallam 1986 for failure through axial splitting

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

and that of Ashby, Sammis 1990 for ductile flow. The results from the splitting model are also used to define the failure strength of individual grains in the new shear faulting model proposed here.

FAILURE U N D E R UNIAXIAL C O M P R E S S I O N
Under uniaxial compression, rock samples break by splitting along the direction of the maximum compressive load. This is induced by the growth of wing cracks originating from the tips of grain boundary starter cracks that undergo shear slip in a frictional manner. The growth of wing cracks has been carefully studied by Horii, Nemat-Nasser 1986 and by Ashby, Hallam 1986. Specifically, Ashby, Hallam 1986 showed that the normalized wing crack length, L=l/ao (see Figure. 1), is related to the applied load in the following way (here and rest of the paper the compressive stresses are assumed to be negative):

K1c
-o', = ~o

(1 + L) 3'2
[1-R-#(I+R)-4.3RL]. 0.23L-~

i ]
Kic =
1.0 MPa~m, crack face friction

(1)

For Westerly granite (see Table 1), the fracture toughness

coefficient ~t = 0.64 and half length of the starter grain boundary crack a 0 = 0.5 mm (Ashby, Hallam 1986), whereas the value of the normalized wing crack length L is not known. It is reasonable to assume that the failure by splitting is induced when the wing cracks reach a critical length large enough to coalesce with their neighbors to cause axial splitting. The experimentally observed failure stress (at zero and very low confinement, where the mechanism of splitting is operable) in granite can be accurately predicted by taking the normalized critical wing crack length to be unity, Figure. 1. Thus, the total extent of the two wings above and below the main crack is approximately equal to one grain size. This is reasonable since splitting occurs through coalescence of such wings along the loading direction. Both the data and model predictions are shown in the biaxial stress space in Figure. 2. The model which is represented by the dashed line closest to the load axes clearly seems to do a good job in predicting whatever limited splitting data is available on granite. Interestingly, when applying the model to other rock types, L =1 turns out to be a problem characteristic constant, independent of the material! It is noteworthy that the failure strength predictions by eqn (1) turn out to be in good accord with the shear faulting model (discussed next) when applied at zero or very low confinements. This can be seen in Figs. 8 to 11 where the predictions of the shear faulting model at low confinements agrees with the dashed line representing eqn (1). Thus, the failure via shear faulting appears to be the rather fundamental from which the splitting failure falls out naturally.

THE SHEAR FAULTING M O D E L


Overall Structure of the Model

To model the above sequence of events, the microstructure of the specimen was divided into a discrete number of elements, with each element corresponding to one grain. The elements were assumed to be aligned in independentrows (Figure. 3) such that the problem of fault propagation becomes one dimensional. This approach is justified since the experimental observations suggest the inhomogeneities to grow mainly longitudinally in a direction aligned with the macroscopic shear fault. To model the

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

progressive growth of damage, the increased probability of failure in the neighborhood of already failed element is considered, and quantified by using the concept of stress enhancement factors K r These factors consider the increased probability for failure of the neighboring elements as though due to a K-fold uniform rise in the applied load z'in the affected region. For the case of r adjacent fractured elements the K r factors are defined as (Argon 1974)

z(K, " %'.): -~! Z( T(r))dr,

1 d

(2)

where d is the width of one element (taken equal to the grain size), Z(Z') is the strength distribution, and z'~ is the background shear stress for the case when no element is broken, and related to the applied compressive stress
(Yl

and lateral confinement R through

z**-- 2

-o] (1- R)sin20,

(3)

where ~) is the shear faulting angle defined in the inset of Figure. 6. Since the factors Kr involve a surface averaging of the enhanced stress, with the strength distribution as a weighting function, they were called stress enhancement factors by Scop, Argon 1967 to distinguish them from the ordinary stress concentration factors. By assuming that the cumulative strength distribution can be characterized by a two parameter Weibull distribution, Z(Z') = Cr m, it can be seen that
=

C(%)" K; ,

(4)

which when combined with eqn (2) yields an expression for K r as

K,:[l[(~(r)] dr]
Ld ot

Ilm

) J

(5)

Equation (5) shows the stress enhancement factors to depend only on the exponent m, the element width d, and the shear stress z(r), but not on the scale factor C. Thus, to predict macroscopic failure by shear faulting, only two functional dependencies need to be determined; the stress enhancement factors and the strength distribution of the material. Calculations of these functions are detailed next. Shear Stress Created by a n Elliptical Inhomogeneity o f Reduced Shear M o d u l u s For the purposes of calculating the shear stress concentration, the localized zone consisting of damaged grains is viewed as an elliptical-shaped inhomogeneity with isotropic properties having a reduced shear modulus G* compared with G of the material on the outside. The solution to this elasticity problem is rather straightforward and can be obtained by using the complex variable approach of Muskhelishvili 1953. The solution details are summarized in Gupta, Bergstr6m 1997. Besides the shear moduli G and G* (expressed in terms of E e, E i, v e and vi) the stress concentration depends on the semi-minor and semi-major axes a and b of the ellipse, and the inclination ~) of the axis a with respect to applied compressive stress cy1. The parameters a, b and ~) are kept free in the model, and their values are

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

determined by minimizing the faulting stress ~1. This leads directly to the length (2a) and inclination (~)) of the critical fault nucleus, in addition to the desired compressive strength on the sample.
(Yl

for a given confinement R

The computation of stress concentration requires the elastic properties of the cracked solid inside the inhomogeneity. Following the procedure outlined in Sammis, Ashby 1986, the modulus E i of the cracked solid idealized as an array ofmicrocracks of length 2a 1 and spacing 2h 1 (inset of Figure. 4) can be expressed in terms of the modulus (Ee) and Poisson's ratio (Ve) of the uncracked solid as (Gupta, Bergstr6m 1997)

ei=

e,
l+V.(k,A)z 1+ 2

(6)

where k~ - b / d, and

2-k2k3~2;with k2- 2a ~ I ;k3 - ~d,


is an independent dimensionless crack spacing parameter. A value of ~ = 5.0 with k 2 = 1.0 and k 3 = 10.0 was used as the default value in the analysis. The plot of normalized reduced modulus vs. The normalized crack length al/h 1 is shown in Figure. 4.
The Strength Distribution

The strength distribution of the elements is specified through a two parameter (m. C) Weibull-type distribution,

= C'rm,

(7)

where )~(r) is the number of flaws per unit volume which limits the strength to f o r less. If the distribution of flaw sizes and their influence on the residual strength is known, then the strength distribution can be fully specified. Consider first the correlation between the flaw size and the element strength. Consistent with the experimental observations, the failure of an element is taken upon the formation of a main sliding crack-wing crack unit that occurs under uniaxial loading and shown in the inset of Figure. 1. The reduced strength a given flaw size results in is assumed to be given by the same equation as used for the prediction of the uniaxial failure strength; i.e., failure of one element occurs when the normalized wing crack length, L = l/ao, becomes unity. This assumption avoids the rather complicated problem of specifying a failure condition based on the actual microcrack interactions, which now are implicitly incorporated in the parameter L. To get the distribution of flaw sizes, we follow the idea of Steacy, Sammis 1992 who assumed the starter flaws in Westerly granite to be distributed with a fractal dimension D f = 2.6, having a largest flaw radius a o = d/2 = 0.5 mm (Ashby, Sammis 1990). The meaning of this fractal distribution of starter flaws can be interpreted in the following way; if there are N v flaws per unit volume having a length in the range [a0/2, a0], then there will be 3Nv flaws with a length in [a0/4, a0/2], and; 3 n N~ flaws with lengths in the interval [a0/2n+l, a0/2n ]. This relationship is valid only for 0 < n < nmax, where nmax in this case is an

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

unimportant cutoff value. From eqn (1) we see that the element strength scales with the flaw size as / ~a, hence, by denoting the resolved shear strength corresponding to a 0 by z" 0, we get the following relations:

= c

Z(

o) =

"r o

o, N~ + 3N~ - 4N~ - 2 ,,n C ,_,,,

(9)

Z(2Vo) = N~ + 3N. + 9N~ = 13N~ = 2" C .

(10)

By relating eqns (8) and (9), the Weibull exponent m can be determined to be 4.0. Note however, that a value o f m = 3.4 would have resulted had we used eqns (9) and (10). Since larger flaws are more important for the failure by shear faulting, it is reasonable to use m = 4.0. To determine the Weibull coefficient C we need to know both z" 0 and N~. The value of N~ is obtained by relating it to the initial damage D o of the specimen (Ashby, Sammis 1990)

4tr,
Do =

~3
aao ) ,

(1I)

.5-

for which A s h b y , Sammis 1990 obtained a value of 0.01 by fitting their damage mechanics model to the experimental observations on Westerly granite. For an initial flaw size 2a 0 = 1.0 mm, this value o f D 0 yields N v = 5.57 x 107. For a starter flaw size 2a = 1.0 ram, the uniaxial compressive strength
(Yl

for

granite was determined in the above section to be -280 MPa. This corresponds to a maximum resolved shear stress z" 0 of 140 MPa. For the above values of Nv and z'0, a value of C = 1.45 x 10-25 is determined from eqn (8). The average strength of the elements for a two-parameter Weibull distribution (1951) is given by Argon 1974 to be

which in this case, becomes 280 MPa for V = 7.5 x 10-10 m 3. Since the microstructure of the specimen is divided into elements corresponding to individual grains, the volume of each element Vmust therefore be bracketed by the volume of a cube with sides 2a0, and that of a sphere with diameter 2a 0. The volume V of the element was taken as the average of these two extremes.

Stress Transformation
Once the critical shear stress "gcrit corresponding to fault instability is computed, it is converted to the compressive strength
(Yl

by using the relation

"t'~, = '-------tl (1 - R) sin 2 ~ +/.tcr u 2


where

(13)

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

aN = o'i(Reos 2 0 + sin2 )

(I4)

and g and R are respectively the frictional coefficient and the lateral confinement ratio. This directly leads to a failure condition of the form

a, =
where

c, G-R

(15)

- ~'ce, 6't - sin20 + 2 # c o s 2 O' s i n 2 ~ - 2#sin 2 ~ C2 = sin2 + 2/lcos ~ 4~

(16) (17)

It is interesting to note that C 2 depends only on the friction coefficient and the shear faulting angle ~0. For ~0= 30 and ~t = 0.6, the constant C 2 becomes 0.32, which means that the predicted failure stress by shear faulting approaches infinity as R approaches 0.32. This should correspond to the transition of the failure from faulting to ductile flow. Further, the influence of ~crit is only to scale the failure curve, the shape is completely determined by ~t and ~0.

Computer Simulation of the Shear Faulting Process


A computer program based on Monte Carlo simulation was written to simulate the statistical nature of the shear faulting process. First, all the necessary input variables are specified. These include: the elements in one row (N); the number of element rows (M); the grain size (d); the Weibull exponent (m); the confinement ratio (R); the elastic properties of the uncracked region and the Poisson's ratio of the cracked material (E~, v~, vi); the crack spacing parameter (A); the critical stress intensity factor (K:c); the friction coefficient (~t); and finally, the initial damage in the sample (D0). After the input variables to the model are specified, all elements are given a strength which is distributed according to the Weibull parameters m and C. The program then finds the lowest stress at which any element breaks. It then iteratively increases the applied stress, and at each step checks if any element is experiencing a stress level that is higher than its strength. When this happens, the element is marked as broken and the stress enhancement factors affecting the neighboring elements are updated. The stress level at which all elements in one row are broken is taken as the failure stress for that row. The process is then repeated for each row in the numerical sample. The failure stress for the sample is taken as the lowest stress at which any of the rows break. Finally, this critical shear stress is converted back to the coordinate system of the applied load according to eqns (15-17).

Results
Failure in Westerly granite was investigated in detail. Each input variable was given the following default values unless otherwise specified: N = M = 50, d = 1.0 mm, m = 4.0, R = 0.10, E e = 70 GPa, ve =

ISSN 0148-9062

To cite this paper: Int. J. R o c k M e c h . & Min. Sci. 34:3-4, p a p e r No. 112.

C o p y r i g h t 1997 Elsevier Science Ltd

1C

approximation of a typical loading situation. Figure 5 shows the dependence of the failure stress on the confinement ratio R. The numerical data is described well by the dashed curve: -(Yl = 190"106 / (0.3-R). When the same data is plotted in the biaxial stress space (Figure. 2), it compares rather well with the failure data obtained on Westerly granite by different investigators. Similarly, Figs. 6 and 7 which show the dependence of the flee parameters ~) and k 1 o n the failure stress, indicate a critical shear faulting angle and normalized half fault width of 30 and 1.5, respectively. Both of these are in excellent agreement with the experimentally recorded values in granite o f k 1 = 1 to 1.5 (Wong 1982; Shimada 1983) and ~) = 30 (Paterson 1978; Shimada 1983). Among the various parameters, KIC, m and the grain size d were found to have the most influence on the shear faulting process (Gupta, BergstrSm 1997).

FAILURE U N D E R H I G H C O N F I N I N G P R E S S U R E
As the confining pressure is increased, failure via shear faulting becomes increasingly difficult as the shear stress drops sufficiently. Instead, the sample is populated almost uniformly with microcracks and the failure results due to macroscopic bulging. The failure in this regime is outside the scope of the model presented above. This, however, can be included by addressing the nucleation of feather cracks (Friedman, Logan 1970) that occur due to Hertzian-type contact between the grain facets under the influence of high normal stresses. Thus, even though the shear stress becomes rather low, cracking activity in the form of feather cracks continue within the rock volume. This aspect of the mode is not developed yet. In absence of this, we follow the work of Ashby, Sammis 1990 who address this failure via the Von Mises yield condition, even though at room temperature the local failure is predominantly through cataclastic flow (or microscopic brittle mechanisms), viz.,

4:

(18)

If the confining pressure is applied in such a way that (Y2= (Y3,then (24) can be rewritten as (yy = (yl-(Y3or (yy = (Y3-(Yl.In the above, (yy is the yield strength of rock and may be determined from the hardness/-/, using (yy = H/3. As shown in Figure. 2, a (yy value of 2200 MPa reasonably envelops the high confinement data on Westerly granite.

FAILURE OF O T H E R R O C K TYPES
In previous sections, a theoretical framework for addressing the failure of rocks by the mechanisms of splitting, faulting and ductile flow, was developed. When applied to Westerly granite, the model predictions in all three distinct failure modes (illustrated as straight lines on the graph in Figure. 2) were found to be in good agreement with the experimentally observed failure stresses. To test the versatility of the models, their predictions in all the three failure regimes were compared with experimental data collected on a wide variety of rocks, including Aplite, Dunite, Eclogite, Limestone, Marble, Rocksalt, and Sandstone. Because of the page limitation, results for only selected rock types are presented in Figs. 8 to 11. In these figures, the dashed line closest to the axes is due to the splitting model, the solid line corresponds to the ductile flow model (eqn 18), and the intermediate dashed line corresponds to the new shear faulting model developed here. The input parameters used in the analysis are summarized in Table 1. By examining these figures we see that the model predicts the failure stresses surprisingly well. Even

ISSN 0 1 4 8 - 9 0 6 2

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

though the shear faulting angle is expected to vary with the level of confinement, the model prediction in the form of a straight line in Figs. 8-11 is for ~) = 30 . This is reasonable as ~) has a feeble effect on the desired compressive failure strength, Figure. 6.

CONCLUSIONS
A new shear faulting model that considers the progressive growth of damage by bringing in the statistical distribution of weak grain boundaries within the rock mass was developed. The predictions are in good accord with the experimental data on a variety of rocks. Interestingly, the shear faulting model predicts the low-confinement splitting failure stress rather well, thus indicating the failure via faulting to be the most fundamental. A detailed discussion on the results and model procedure can be found in Gupta, Bergstr6m 1997.

ACKNOWLEDGMENTS
This work was supported by the ONR Grant No. N00014-92-J-1247 and ARO grant No. for studying the nucleation of cracks in saline and freshwater ice. For this we are grateful to Drs. Y. D. S. Rajapakse, Steve Ramberg and Tom Curtin of that agency.

FIGURES

ISSN 0148-9062

To cite this paper: Int. J. R o c k M e c h . & M m . Sci. 34:3-4, p a p e r No. 112.

C o p y r i g h t 1997 Elsevier Science Ltd

Paper 112, Figure 1.


30O0
II J, . .

ranlte

,,=

t~ I
C~

2500

[] .,.

SpJitl.ing Prediction Mogi (19661 Sc'h, oc~ Heard (1974) ShJmada (1981)

i-,

-,--___.

c~
[]

t 6o
.J $-4

.9

.~ 15oo if?
",/3 c.) e-,
+

looo
5O0
N

/
[

~3
6
~L

40
02 <9"2

,4

"~ N

=I~

20
Z
J

Q
0

.,1
|

0.1

0-2

0,3

C o n f i n e m e n t Ratio, R
Figure 1.

ISSN 0 1 4 8 - 9 0 6 2

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 2.


A x i a l Stress, o l [MPa]

E-"
x

-5000
0

-4000 L

-3000 / z z7
!

-2000
|

-1000
J -Fir B

0
0

._~nl.~__.

- ~----=-*'~
_.1~..~- ~ ~' "

-20
---.

............

s--T

_._ - - ' D " - "

. ,.#'

-1000

-2000
-40
iJ IL Q ..' .~ .

-3000

-60

.,k
[]

L[.: ]

, ,."

-4000

-80

o
--

I~og, i (1966)

Schoc~k Hea=d (174)


Shimacla (1981)
Axial , ~ l ~ i n ?

~
./
j

.a.9./

,," /,' ,/
," ."

.:'

[ ,! ["
:'
i

! -5Oo0

Shear Faultirlg
Ouotile Flow

.~"
-/
r -

-180 1130

.6000
0

-.','~

Not~a.tized Axial Stress. ~ t / ( K

10 -~)

F i g u r e 2.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 3.

Ro t

Figure 3.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 4.


i I | i i , i i i i i i i i i i i i

"-.

r-C.

--

~f

'

0.01
,%,

-iIi , i
--.T 'L . . . . . .

F~ z
0.00l OA
1

Io

Normalized Cr~k Length,


Figure 4.

at~hi

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 5.


,,.i r, i

Grani~
2000
c' c"

,"
+1' .i

~=19~(.0,3-R1

. . . . .

---__

,i
70

!
6O

~, I
,~,
_,

1500
j"

50 40
30
n:J

1000
$ .-...+"4

4 .," .+-"

,e4

soo

20

.N
el

(.J

-..=

0 0.4

7~

0.1

0.2

0,3

O o n f m e m e n t R~tio, R

Figure 5.

ISSN 0148-9062

To cite this paper: Int. J. R o c k M e c h . & M m . Sci. 34:3-4, p a p e r No. 112.

C o p y r i g h t 1997 Elsevier Science Ltd

Paper 112, Figure 6.


2OOO 1800
I I I . I I

Granite

,~
4

+.
70
"t

,y

1600
1400

60

E,
I+ * % ,6" + + % ,t I-~

I
I,..,

1200 10I~

50
-+

40
t- %
~1

I-

800L-, "..%

.m

-"

'

30

6O(> 4130

cZ
5~

"72

%
10
I

2O0 O
L

E L.=
Z

.......

l ....

0 60

10

20

30

40

5O

Shear Faulting Angle,


F i g u r e 6.

ISSN 0 1 4 8 - 9 0 6 2

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 7.


20O0
I I I

G ra rlEte 70
+ -I-

1800 1600
P_.., 1400

-----

E
5o
4 14

-r200
1000 8O0

p2
-o3=
El

f
h-,

40 30 2O

~j
6OO r.,.
,:d

,~ L~

"U E3
.4

40O 200
0
I I r

2".

Normalized Shear l,'~ulting Width,


F i g u r e 7.

bid

ISSN 0148-9062

To cite this paper: Int. J. R o c k M e c h . & M m . Sci. 34:3-4, p a p e r No. 112.

C o p y r i g h t 1997 Elsevier Science Ltd

P a p e r 112, Figure 8. Axial St,r::,~s,~, [MPa]


.6000 0
I

-500,0
I

-40"130
I .

-3DO,O
_ I

-213'00
I

-1000
I .._ .~ 4

__;~ibbro

...............

..................

-_./ ...........

....J

;..:- ......

-I/ :2'+/:l
-100'0

-20

C
..r,.
,.2-

..._._..._--'''"'"

-200"0

-40

-3000
s~

-,6[3

':f2 ,4000

o -q o Z

-BO

_-50..30

-fO0

-6000
-t'130
-50

Normalized Axial Stress: al,/{K x 10-'~}


Figure 8.

ISSN 0 1 4 8 - 9 0 6 2

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 9.


Axial S*,re_s~, ol [MPa]
-8000
I

-7000
:I

-6000
L

.5000
J .....

-4,OOO -3O00
-/-:
I ......

-2000
~" - : -~;-' ~.= ...--

-1000
l ..... I

O
13

e
-10 -20

.E~!~gi~ . . . . . . . . . . . . . .
~ _ _ _ _ .

....... --- "''''~'""

- """ +

". . . . . . "

C -30 u4 ~/
r-~

-40
-60" 50

.~ ""~ "1 ,' : ,'J


~' .iI.l"l

, ?"

-1000 -2000 -3000


-4000

,%-

/ " " '


---F

.!:/

-50g0
-6000

.E('?
%

o~

,...o

-70
-8(1 -[

SNm~da{ 19~,}
Axial S_pli~ling Sh, ear t-aulSn 9 OUI:iEn Flow
c i

_ _ -80

../
i i

/-,

.
i

,.

,,
i |

=7000
-13000 0 -2.0 -1 l?

-70

-60

-50

-40

-30

N o r m a l i z e d Axial

Stress= eE/(K i0 -2)

Figure 9.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 10.


Axial Stress, cy, [MPa]
N"
O

.70OO 0

6o00

-N300 -4803

-3o00

-2000

.1003

-Ip~._~.e '_2.'._. '- ...... 2._~-...J __ 7-j~.~._;:~..':: ....


...... -~..... :'"~ .......

N L<
--...

.,4. - t 0 0 0
i":!]
.2000

-10

-A
E.O

-20

-30
"-C .-i

'

r,-t

-40 ..qhima~a [1983) ...... ANal $421illin~ ......... Shear Faulting Ouclile Flow
i 1 -.

/~,']'
f

]
i - ~ . - -

-3000

C'
c.~

-4000
6":

-5000

.-=

-50 .,

,/
i

7,

--6000
4- -7000

-50

-40

-30

-20

.10

Normalized Axiat Stress,

o~/(K

x 10 -~)

F i g u r e 10.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Figure 11.


Axial Stress, ot [_MPa]
-soo
I

-,0o
I.__ .....

-3oo
I ___

-2oo
.I

-1 o0
[ ._~_a

5ar~lelo.r~-

.....

. ~-,~--'"

.....

-0.5
.__-'~F ._--"

, ..... ~--..~'" .--~.....


. .i-~ "'~" - -

-50
-100
'4,,"

---__

-1

ra
~ -1.5
," t

.t5O
.20qO

o3
..a .22; =-~

-2
-2.5

." t
.'4

-250 -300

r.r)

,(' ,'

,%
-33,0
,"5

e_,

-3

Gow~l , a ~ R u m m e l { 1 9 8 0 ) A x i a l S plitldng

-44)0
,'

-3.5

........
- -

Shean"Faulting
Ductile Fqow

,," ,,'
-2 ,1

-4~;0 -500

;7.
-8

Normalized Axied S~ret--~, c~l/(K x 10

'~)

F i g u r e 11.

TABLES

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Paper 112, Table 1.


Table 1: Material properties used in the theoretical model. Material
2o.o Kic # E~
u,,

= ul

Do

oy

[mml [MPav~
(i) Aplite Dunite Eclogite Gabbro Granite Limestone Marble Rocksalt Sandstone (~) 0.2 1.0 0.6 0.8 1.0 0.05 0.4 0.5 2.0 (s) 1.0 1.0 1.0 1.0 1.0 0.4 0.2 0.2 0.3
0.6 0.6 0.6 0.55 0.64 0.55 0.6 0.55 0.6

[GPa]
(0
70 150 130 92 70 77 70 37 130 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.3 0.2 0.08 0.04 0.08 0.I2 0.01 0.I5 0.12 0.20 0.I5

[GPa] [MPa]
(5) 5 5 5 5 5 5 5 5 5 (6) 4 4 4 4 4 4 4 4 4 (~) . . . . . 58 2300 i30 3500 94 4000 60 2200 58 2200 70 700 70 350 37 65 130 1000

References
References
Argon A. S. 1974. Composite Materials, Statistical Aspects of Fracture, 5, pp. 153-189, New York, Academic Press. Argon A. S. et al. 1983. Fracture in compression of brittle solids. Technical Report NMAB-404, National Academy Press. Ashby M. F., Hallam S. D. 1986. The fracture of brittle solids containing small cracks under compressive stress states. Acta metalI., 34:3, 497-510. Ashby M. F., Sammis C. G. 1990. The damage mechanics of brittle solids in compression. Pure andApplied Geophysics. 133:3, 489-521. Bailey C. D., Hamilton Jr. J. M., Pless W. M. 1979. Mate~ EvaI., 37:6, 43. Batzle M, Simmons G., Siefgried R. 1979. Direct observation of fracture closure in rocks under stress. Eos Trans., 60, 380. Brace W. G., Paulding B. W., Scholz C. 1966. Dilatancy in fracture of crystalline rocks. J. Geophys. Res., 71:16, 3939-3953. Fairhurst C., Cook N. G. W. 1966. The phenomenon of rock splitting parallel to the direction of maximum compression in the neighborhood of a surface. In Proc. 1st. Congn Int. Soc. Rock. Mech. Lisbon, 1, pp. 215-232. Friedman M., Logan J. M. 1970. Microscopic feather fractures. Geological Society of America Bulletin, 81, 3417-3420. Gowd T. N., Rummel F. 1980. Effect of confining pressure on the fracture behavior of a porous rock. Int. J. Rock. Mech. Min. Sci. and Geomech. Abst~, 17. 225-229.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Min. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Griggs D., Handin J. 1960. Observations on fracture and a hypothesis of earthquakes. In Rock deJormation (ed. D. Griggs and J. Handin), Mere. Geol. Soc. Am., 79, pp. 347-364. Gupta V., Bergstrom J. 1997. Compressive failure of rocks via shear faulting, in preparation. Horii H, Nemat-Nasser S. 1985. Compression-induced microcrack growth in brittle solids: axial splitting and shear failure. J. Geophys. Res., 90:B4, 3105-3125. Horii H, Nemat-Nasser S. 1986. Brittle failure in compression: splitting, faulting and brittle-ductile transition. Phil, Trans. R. Soc. Lond., 319, 337-374. Kachanov M. L. 1982. A microcrack model of rock inelasticity, part I, Frictional sliding on microcracks. Mech. Mat., 1, pp. 19-27. Kranz R. L 1983. Microcracks in rocks: a review. Tectonophysics, 100, 449-480. Mogi K. 1966. Some precise measurements of fracture strength of rocks under uniform compressive stress. Rock Mech. Eng. Geol., 4, 41-55. Muskhelishvili N. I. 1953. Some basic probIems of the mathematical theory of elasticity. P. NoordhoffLtd. Nemat-Nasser S., Horii H. 1982. Compression-induced nonplanar crack extension with application to splitting, exfoliation, and rockburst. J. Geophys. Res., 87:B8, 6805-6821. Nemat-Nasser S., Obata M. 1988. A microcrack model of dilatancy in brittle materials. J. Appl, Mech., 55, 24-35. Paterson M. S. 1978. Experimental Rock Deformation--The Brittle Field, Springer-Verlag, New York. Rothman R. L. 1977. Proc. of the first conference on acoustic emission/microseismic activity in geological structures and material, p. 87, Edited by H. R. Handy Jr., and F. W. Leighton, Clausthal FRG: Trans. Rudnicki J. W., Rice J. R. 1975. Conditions for the localization of deformation on pressure-sensitive dilatant materials," J. Mech. Phys. Solids', 23, 371-394. Scop P. M., Argon A. S. 1967. Statistical theory of strength of laminated composites. J. Composite Materials' 1, 92-99. Schock R. N., Heard H. C. 1974. Static mechanical properties and shock loading response of Granite. J. Geophys. Res., 79:11, 1662-1666. Scholz C.H. 1968. Microfracturing and inelastic deformation of rock in compression. J. Geophys. Res., 73, 1417-1432. Shimada M. 1981. The method of compression test under high pressures in a cubic press and the strength of Granite," Tectonophysics, 72, 343-357. Shimada M., Cho A, Yukutake H. 1983. Fracture strength of dry silicate rocks at high confining pressures and activity of acoustic emission. Tectonophysics, 96, 159-172. Shimada M. 1986. Mechanism of deformation in dry porous basalt at high pressures. Tectonophysics, 121, 153-173. Steacy S. J., Sammis C. G. 1992. A damage mechanics model for fault zone friction. Journal of Geophysical Research, 97: B1, 587-594. Tada H. P., Paris R C., Irwin G. R. 1985. The stress analysis of cracks handbook. Del Res., St. Louis, Mo. Tapponier R, Brace W. F. 1976. Development of stress induced microcracks in Westerly granite. Int. J. Rock. Mech. Min. Sci. & Geomech. Abstr, 13, 102-112. Tullis J., Yund R. A. 1977. Experimental deformation of dry Westerly granite. J. Geophys. Res., 82, 5705-5718.

ISSN 0148-9062

To cite this paper: Int. J. Rock Mech. & Mm. Sci. 34:3-4, paper No. 112.

Copyright 1997 Elsevier Science Ltd

Weibull W. 1951. A statistical distribution function of wide applicability. J. Appl. Mech., 18, 293-297. Wong T.-T. 1982. Micromechanics of faulting in Westerly granite. Int. J. RockMech. Min. Sci & Geomech. Abstn, 19, 49-64.

ISSN 0148-9062

You might also like