You are on page 1of 13

International Journal of Engineering Science 41 (2003) 917929

www.elsevier.com/locate/ijengsci

Anisotropic dynamic damage and fragmentation


of rock materials under explosive loading
a,*
Yong-Qiang Zhang , Hong Hao b, Yong Lu a

a
Protective Technology Research Center, School of Civil and Environmental Engineering,
Nanyang Technological University, Singapore 639798
b
Department of Civil and Resource Engineering, University of Western Australia Nedlands, WA 6009, Australia
Received 16 September 2002; accepted 11 October 2002

Abstract
This paper describes the development of a constitutive model for predicting dynamic anisotropic damage
and fragmentation of rock materials under blast loading. In order to take account of the anisotropy of
damage, a second rank symmetric damage tensor is introduced in the present model. Based on the me-
chanics of microcrack nucleation, growth and coalescence, the evolution of damage is formulated. The
model provides a quantitative method to estimate the fragment distribution and fragment size generated by
crack coalescence in the dynamic fragmentation process. It takes account of the experimental facts that a
brittle rock material does not fail if the applied stress is lower than its static strength and certain time
duration is needed for fracture to take place when it is subjected to a stress higher than its static strength.
Numerical results are compared with those from independent eld tests.
2003 Elsevier Science Ltd. All rights reserved.

Keywords: Anisotropic damage; Fragmentation; Second rank damage tensor; Rock material

1. Introduction

It is well known that rock fragmentation by blasting is a dynamic fracture process. In rock
blasting, it is generally understood that the stress waves have signicant contribution to damage
and fragmentation. The propagation, reection and interaction of the stress waves result in
crushing, spalling and fragmentation of rock materials.

*
Corresponding author. Tel.: +65-67906199; fax: +65-67910676.
E-mail address: cyqzhang@ntu.edu.sg (Y.-Q. Zhang).

0020-7225/03/$ - see front matter 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0020-7225(02)00378-6
918 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

Due to the extreme complexities of the phenomena in dynamic fracture and fragmentation, the
majority of blasting models today are based on empirical or semi-empirical formulas. They tend
to overlook the physical laws governing the processes in blasting. Blasting models based on
constitutive relationships are of special interest because of their capabilities in describing the main
features of fracture and fragmentation. With the advance of high-speed computer methods, eorts
have been directed toward developing continuum descriptions of fracture, fragmentation, and
wave propagation to evaluate complex fracturing events. Shockey et al. [1] have pursued models
based on the activation, growth, and coalescence of inherent distributions of fracture-producing
aws, predicting crack and fragment size spectra resulting from explosive rock fracture. In recent
years, a number of microcrack-based constitutive blasting models have been published [27].
Grady and Kipp [2] presented a description of dynamic fracture and fragmentation of rock that
emphasizes the strain-rate dependence of measurable fracture properties such as fracture strength,
fracture energy and fragment size. Taylor et al. [3] developed a damage model to simulate stress
wave induced rock fracture during blasting based on the analysis of cracked systems [8,9].
In some early continuum models [24], it is assumed that microcracks initiate and grow im-
mediately when the strain becomes tensile. However, it is observed from the experimental results
that rock materials do not fail at tensile stresses below their static tensile strength. Furthermore,
Bawden et al. [7] pointed out that, in dynamic loading, the tensile stress that exceeds the material
strength may not damage the material if its duration is too short. Blast damage is accumulated as
a function of time and applied stress. The fact is supported by the recent experimental work of Li
et al. [10]. In the dynamic damage models proposed by Liu and Katsabanis [5] and Hao et al. [6], a
critical strain value and certain time duration are introduced to model the damage initiation and
accumulation.
It is observed that the cracks by blasting are strongly oriented. The inuences of the shape, size,
orientation and distribution of cracks in a rock mass, which usually result in dierent material
properties in dierent directions owing to their usual predominant orientation in certain direction,
cannot be captured by an isotropic damage scalar. To this end, an anisotropic damage model is
presented for dynamic damage of initially isotropic rock materials under blasting loading in this
study. The description of anisotropic damage behavior of rock mass under explosive loading is
developed by employing a second rank symmetric damage tensor [11] in the irreversible ther-
modynamic frameworks. The model provides a quantitative method to estimate the fragment
distribution and fragment size generated by crack coalescence in the dynamic fragmentation
process. Numerical results are compared with those from independent eld tests.

2. Damage variables

Under applied stresses, microcracks can nucleate and grow. These microcracks are generally
distributed in some preferential orientations. Consequently, the growth and nucleation of mi-
crocracks not only result in the reduction of elastic moduli, but also in anisotropic behaviour. For
most initially isotropic rocks, induced anisotropy can reasonably be simplied in the orthotropic
case produced by three orthogonal principal sets of microcracks. Damage in rock materials is
usually induced by nucleation, growth and coalescence of microcracks. Since the development
of these microcracks is governed by the action of applied stress and strain, material damage is
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 919

essentially anisotropic. This feature is especially important in rock materials damaged by the
development of distributed and oriented microscopic cracks [12].
As scalar damage variables often have serious limitation to the description of actual material
damage, a number of theories have been developed to model the anisotropic damage state by
means of damage variables ranging from a vector to higher rank tensors. Vector damage variables
[1316] can be applied to describe the distributed plane cracks by describing the normal direction
of each plane crack and the crack density of the specic plane. However, the vector damage
variables cannot represent the eect of a set of plane cracks of dierent orientations by means of a
simple addition of the corresponding vectors. Moreover, odd rank tensors in general can be ex-
cluded from the set of internal variables, since they do not conform to the invariance with respect
to the rotation of the frame [17]. Consequently, in spite of the complexity of their mathematical
structures, more elaborate theories have been proposed for the description of the state and the
eect of the distributed cavities by use of higher and even rank tensors. Among them, second rank
symmetric damage tensors [11,17,18] are most commonly employed because they are mathe-
matically simpler than the higher rank tensors, and yet can describe most essential features of
anisotropic damage.
To develop a feasible damage theory in this paper, it is postulated that the anisotropic damage
state can be described by a damage tensor

X
3
D Di ni  ni 1
i1

where Di and ni are the principal value and the unit vector of principal direction of the tensor D.
The damage tensor D of Eq. (1) was derived by Murakami [11] by postulating that the principle
eect of the material damage consists in the net area reduction due to three-dimensional distri-
bution of microscopic cavities in the material. Hence, Di in Eq. (1) can be interpreted as the ratio
of area reduction in the plane perpendicular to ni caused by the development of microcracks.
Though the second rank symmetric damage tensor D cannot describe more complicated damage
state than orthotropy, it has been often employed in the development of anisotropic damage
theories [11,1921].

3. Constitutive equations

A rock material usually suers damage by the development of distributed microscopic cracks
and leads to the nal fracture by their coalescence. In order to represent the state of anisotropic
damage characterized by these cracks, a second rank symmetric damage tensor D dened in
Section 2 will be employed.
As a point of departure, we postulate that the total strain tensor e can be decomposed into
elastic and plastic parts

e ee ep 2

and the free energy function can be expressed as [21]


920 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

wee ; q; C wed ee ; C wpd q; C 12ee : C : ee wpd q; C 3

where C is the damaged elasticity tensor, q denotes a suitable set of plastic variables, and ac-
cording to Clausius Duhem inequality, it has
   
e 1 e _ e owpd _ e p
owpd
e_ : r  C : e  e :C:e : C e : C : e_   q_ P 0 4
2 oC oq

Consequently, the stressstrain relation can be obtained as follows:

ow
r C : ee 5
oe

For the damage-dependent elasticity tensor C, a great many expressions have been proposed [22
24]. By introducing an intermediate second-order tensor U, Swoboda [25] dened the elasticity
tensor as

C kU  U 2lUU 6

where the symbol  denotes the dyadic product of two tensors, the symbol  denotes the
symmetrized dyadic product dened as AB : C A  C CT =2  BT for any arbitrary second
order tensors A, B, C, and

U A1 I A2 D A3 D  D 7

where A1 , A2 and A3 are functions of the invariants of D, k and l are Lames constants of virgin
material, and I is the second-order identity tensor. As the elasticity tensor must satisfy the positive
denite condition, it requires the elastic potential function

1 k
wed ee ; C ee : C : ee U : e2 le  U : U  e P 0 8
2 2

Note that e  U : U  e P 0 if U is a positive denite tensor. Thus, a fourth-order positive


denite problem reduced to a second-order one.
Although not necessarily, the normalized damage variable is convenient to use. Here it is
postulated that

C0 if D 0
C 9
0 if D I

where C0 is the isotropic elasticity tensor of the virgin material, which can be expressed as

C0 kI  I 2lII 10
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 921

Based on this condition, we have


A1 1; A1 A2 A3 0 11
Assuming that A1 , A2 and A3 are material constants and noting the positive denite requirement
on U, we know that
A1 1; A2 k; A3 1  k 06k61 12
Hence, there is only one independent material constant k. Under the hypothesis of complementary
energy equivalence, it has k 1. Then the elasticity tensor takes the form
C I  D  C0  I  D 13
Substituting Eqs. (1) and (10) into Eq. (13) yields
ri 1  Di kh  n 2l1  Di ei i 1; 2; 3; no summation 14
where

h e1 e2 e3 ; n D1 e1 D2 e2 D3 e3

4. Damage evolution

Under high-rate loading, microcracks will be initiated and grow in a rock mass, and damage
will be accumulated. As a certain time duration is needed for fracture to take place when a rock
material is subjected to a stress higher than its static strength, according to the denition of the
damage variable, the evolution of damage in the ith principal direction (i 1, 2, 3) can be de-
termined by the number of cracks which activate at the time t as follows:
Z t
Di t N_ i sAt  s ds 15
tci

where tci is the time duration needed for fracture to take place, and At  s is the stress-relieved
area which is determined by a microstructural law for the growth of cracks activated at past time
s. Assuming the microcracks are penny shaped, it has

At  s pc2g t  s2 16

where cg is
pthe
crack
growth velocity, and by studying the dynamic crack propagation, the relation
cg 0:38 E=q is usually assumed [26]. It should be noted that the derivation of Eq. (15) is based
on the assumption that the growth velocity reaches cg very quickly as soon as a crack activates.
As for the variable Ni in Eq. (15), it is the number of microcracks per unit area in the plane
perpendicular to the ith direction, and the rate of microcrack activation can be calculated by

b
N_ i ahei  ecr i 17
922 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

which is obtained by extending that dened by Yang et al. [27]. In Eq. (17), the angular bracket hi
denotes a function dened by hxi jxj x=2, and in which a and b are material parameters, and
ei is the principal strain in the ith direction. Substituting Eqs. (17) and (16) into Eq. (15) gives
Z t
Di t apcg hei  ecr ib t  s2 ds
2
18
tci

5. Fragmentation

Fragments are associated with crack initiation, propagation and coalescence, thus it is neces-
sary to know the crack size in order to predict the fragment size. For this reason, the damage
dened by Eq. (15) is given in terms of the distribution of crack size r by setting r cg t  s,
Z cg ttci
Di t xi r; t dr 19
0

where
pr2 _
xi r; t Ni s 20
cg

is the damage distribution in the ith direction, in which the variable s t  r=cg .
Full fragmentation in the ith direction is dened to occur when the damage

Di tfi 1 21

which corresponds to fracture coalescence at time tfi . At fracture coalescence, it is assumed that
the fragment sides are formed by the fracture faces. Therefore, at time tfi there is a correspondence
between the fragment size distribution function and the damage distribution function in the ith
direction

F Li 12xLi =2; tfi 22

where Li is the fragment size in the ith direction corresponding to crack radius r Li =2.

6. Identication of the parameters

In the present damage and fragmentation model, four parameters, namely a, b, cg and ecr , need
to be determined from the dynamic fracture properties of rock materials. Since most rate-
dependent rock tests performed are uniaxial, the static failure strain ecr can be easily determined
from uniaxial static tensile test results,
rst
ecr 23
E
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 923

where rst is the static tensile strength and E is Youngs modulus for intact material. The crack
growth velocity cg can be obtained from the relation equation shown in Section 5. The other two
parameters can be determined by using results obtained from uniaxial tensile test. This is dis-
cussed in the following.
Assuming the uniaxial strain rate is constant, the tensile strain can be expressed as

e e_0 t 24

where e_0 is the constant strain rate of uniaxial tension, and the axial damage can be derived from
Eq. (18)

Dt me_b0 ht  tc ib3 25

where the relation tc ecr =e_0 is used, and

2apc2g
m 26
b 1b 2b 3

is a constant, depending on the material properties.


In the case of uniaxial tension, a rock material is damaged by the development of distributed
microscopic cracks and leads to the nal fracture by their coalescence without signicant inelastic
deformation [28]. Thus for this case, ignoring the plastic strain, namely e ee , the axial stress can
be obtained from Eq. (14) as follows:

r 1  D2 Ee 27

where the relation E l3k 2l=k l is used.


If the tensile strain and damage scalar corresponding to the fracture stress rF are denoted by eF
and DF , respectively, from Eqs. (27) and (24), it has

rF 1  DF 2 EeF and eF e_0 tF 28

where tF is the total time to reach the fracture stress. From Eq. (25), it has
 1=b3
DF b=b3
tF  tc e_0 29
m

Combining Eqs. (28) and (29), and using the relation ecr e_0 tc , the fracture stress at a certain
strain rate in uniaxial tensile can be obtained as
 1=b3
2 2 DF 3=b3
rF 1  DF rst E1  DF e_0 30
m
924 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

Dependence of the fracture stress on strain rate is provided by the above equation. Since
fracture stress for many brittle materials such as rock and concrete depends on the cube root of
the strain rate [5,2931], b can be taken as equal to 6.
It is evident that the fragment size distribution is also dependent on strain rate. As for the
fragment size predictions for the case of uniaxial tension with a constant strain rate, from Eq. (20)
it has
apr2 b
xr; t e_ ht  tc  r=cg ib 31
cg 0

Fragmentation occurs at a time tf corresponding to Dtf 1. From Eq. (25), it has


b=b3
tf m1=b3 e_0 tc 32

The axial fragment size distribution, Eq. (22), can be maximized with respect to the axial frag-
mentation size L. Specialized to a uniaxial tension and a constant strain rate the axial fragment
distribution is

apL2 b
F L e_0 tf  L=2cg  tc b 33
8cg

This has a maximum for


4cg
Lm tf  tc 34
b2
Substituting Eq. (32) into the above equation yields
4cg 1=b3 b=b3
Lm m e_0 35
b2
This is the expression for the dominant fragment size (fragment size corresponding to the largest
volume fraction of material) that is dependent on the strain rate.
The parameter a can be determined by using the predicted dependence of fracture stress on
strain rate from Eq. (30) and dominant fragment size on strain rate from Eq. (35) to obtain a best
t with the data obtained by experiments. According to the numerical investigations and some test
results of rock materials under explosive loading [27,32,33], the damage value is about 0.22 when
the dynamic tensile stress reaches the dynamic failure stress, namely DF 0:22, which is also the
minimum damage value for the beginning of fragmentation.

7. Application

7.1. One-dimensional loading

In this section, the response of oil shale subjected to a tensile stress is studied to verify the above
theoretical derivations. The oil shale with kerogen content approximately 80 ml/kg is used for the
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 925

analysis. The representative properties of 80 ml/kg oil shale are: elastic modulus E 17:8 GPa,
yield strength rs 50 MPa, density q 2260 kg/m3 , Poissons ratio m 0:27, elastic wave speed
of cl 2:8 km/s, and static tensile strength rst 5 MPa.
Using the above oil shale properties and the test data reported by Grady and Kipp [2], the
material parameters in the model will be determined for oil shale. The parameter b is taken to be
equal to 6 so that the fracture stress is cube root dependent on the loading rate, and the crack
growth velocity cg is 1066 m/s that is calculated from the above relation expression. The parameters
a is taken to be a 1:2 1025 /m2 s that is determined by tting Eqs. (30) and (35) to the test data
shown in Figs. 1 and 2.
Using these parameters and Eqs. (30) and (35), the fracture stress and dominant fragment size
as functions of strain rate are estimated. The corresponding curves are also shown in Figs. 1 and
2. As can be seen, the predicted values of the fracture stress and the dominant fragment sizes agree
reasonably well with the test data.

Fig. 1. Fracture stress dependence on strain rate for oil shale.

Fig. 2. Fragment size dependence on strain rate for oil shale.


926 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

Fig. 3. Fragment distributions corresponding to dierent constant strain rates.

Fragment distributions calculated from Eq. (33) for the three constant strain rates are shown in
Fig. 3. As can be seen, the fragment sizes at the strain rate of 104 s1 are very small with a
dominant size of about 0.77 mm. On the other hand, at lower strain rates of 103 s1 and 102 s1 the
dominant fragment sizes are about 3.6 and 17 mm, respectively.

7.2. Axisymmetric simulation of blasting crater

A single borehole blasting experiment was conducted in the nominal 80 ml/kg oil shale [33,34].
The blasthole was perpendicular to the ground surface with a diameter of 0.162 m. The explosive
column had a length of 2.5 m and the stemming length was 2.5 m. The detonation point was
positioned at the bottom of the explosive column and on the line of symmetry, and the explosive
used in the test was IREGEL 1175 U. In this section, the crater created from this test is nu-
merically simulated to further verify the above theoretical derivation.
The present damage and fragmentation model is implemented into the commercial program
AUTODYN2D [35] as its user subroutine. The model parameters have been determined at
Section 7.1. Fig. 4 shows the conguration of the numerical model, in which explosive was
simulated by Euler processor and was modeled using the JonesWilkensLee (JWL) equation of
state in AUTODYN2D. The oil shale was assumed to satisfy the linear EOS and simulated by
Lagrange processor. The Mohr-Coulomb yield criterion is used to calculate the plastic ow of the
oil shale. The whole domain is assumed to be axial symmetric. The transmitting boundary
technique is used in AUTODYN to reduce reection of shock wave from the specied bound-
aries. These transmitting boundaries allow outward traveling waves to pass through without
reecting energy back into the computational grid. Only the normal component of velocity of the
wave is dealt with and the velocity component parallel to the boundary is assumed to be unaf-
fected by the boundary.
The JWL equation of state models the pressure generated by chemical energy in an explosive. It
can be written in the form
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 927

Fig. 4. Conguration of numerical modeling.

   
x r1 V x xw
P C1 1 e C2 1  er2 V 36
r1 V r2 V V

where C1 , C2 , r1 , r2 and x are constants. P is the pressure, V is the relative volume q0 =q, and q0
and q are the initial density and the current density respectively. w denotes the internal energy.
The parameters in JWL equation of state for IREGEL 1175 U used in the present study are listed
in Table 1, in which w0 is initial ChapmanJouguet (CJ) energy per volume as the total chemical
energy of the explosive, and VOD is CJ detonation velocity of the explosive.
The region of signicant rock damage in the oil shale is usually measured by excavating the
rock loosened by the blast. The crater formed in the experiment was excavated and its dimensions
surveyed [3]. The radius of the crater is about 4.9 m, the depth of the crater on the line of
symmetry is about 1.75 m. This is compared to the calculated rock damage regions dened by the
contour in the material where the damage scalar X exceeds 0.22. The damage scalar X is dened as
X min Di t, and here we assume that the rock is loosened enough for us to be excavated when
the damage scalar X exceeds 0.22.
The damage zone where the damage scalar X exceeds 0.22 is shown in Fig. 5 at a time close to
terminal damage growth. As can be seen, the predicted radius on the ground of the predicted
crater is approximately 5.02 m which is very close to the experimental value 4.9 m. The predicted
depth of the crater on the line of symmetry is approximately 1.68 m which is rather close to the
experimental result of 1.75 m.

Table 1
JWL parameters used for modelling IREGEL 1175 U in the present study
C1 (GPa) C2 (GPa) r1 r2 x w0 (MJ/m3 ) VOD (m/s) q0 (kg/m3 )
47.6 0.524 3.5 0.9 1.3 4500 6178 1250
928 Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929

Fig. 5. Calculated damage zone where the damage variable X exceeds 0.22.

8. Conclusions

By employing a second rank symmetric damage tensor in the irreversible thermodynamic


frameworks, a model for dynamic anisotropic damage and fragmentation of rock materials under
explosive loading has been developed. It considers the experimental facts that a brittle material
does not fail if the applied stress is lower than its static strength and certain time duration is
needed for fracture to take place when it is subjected to a stress higher than its static strength.
Based on the mechanics of microcrack nucleation, growth and coalescence, the evolution of
damage is formulated. The model provides a quantitative method to estimate the fragment dis-
tribution and fragment size generated by crack coalescence in the dynamic fragmentation process.
The various damage activation parameters involved in the model can be easily determined. The
model has been calibrated by a eld blasting test.

References

[1] D.A. Shockey, D.R. Curran, L. Seaman, J.T. Rosenberg, C.F. Petersen, Fragmentation of rock under dynamic
loads, Int. J. Rock Mech. Min. Sci. 11 (1974) 303317.
[2] D.E. Grady, M.E. Kipp, Continuum modelling of explosive fracture in oil shale, Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. 17 (1980) 147157.
[3] L.M. Taylor, E.P. Chen, J.S. Kuszmaul, Micro-crack induced damage accumulation in brittle rock under dynamic
loading, Comput. Meth. Appl. Mech. Engng. 55 (1986) 301320.
[4] J.S. Kuszmaul, A new constitutive model for fragmentation of rock under dynamic loading, in: 2nd International
Symposium on Rock Fragmentation by Blasting, Keystone, Colorado, 1987, pp. 412423.
[5] L. Liu, P.D. Katsabanis, Development of a continuum damage model for blasting analysis, Int. J. Rock. Mech.
Min. Sci. 34 (2) (1997) 217231.
[6] H. Hao, G.W. Ma, Y.X. Zhou, Numerical simulation of underground explosions, Int. J. Blast. Fragment. 2 (1998)
383395.
Y.-Q. Zhang et al. / International Journal of Engineering Science 41 (2003) 917929 929

[7] W.F. Bawden, P. Katsabanis, R. Yang, Blast damage study by measurement and numerical modelling of blast
damage and vibration in the area adjacent to blast hole, in: W.F. Bawden, J.F. Archibald (Eds.), Innovative Mine
Design for the 21st Century, Proceeding of the International Congress on Mine Design, Kingston, Ont., Canada,
1993, pp. 2326.
[8] B. Budiansky, R.J. OConnell, Elastic moduli of a cracked system, Int. J. Solids Struct. 12 (1976) 8197.
[9] L.G. Margolin, Elasticity moduli of a cracked system, Int. J. Fract. 22 (1983) 6579.
[10] X.B. Li, S.R. Chen, D.S. Gu, Dynamic strength of rock under impulse loads with dierent stress waveforms and
durations, J. CentralSouth Inst. Min. Metall. 25 (3) (1994) (in Chinese).
[11] S. Murakami, Mechanical modeling of material damage, J. Appl. Mech., Trans. ASME 55 (1988) 280286.
[12] J.L. Chaboche, Continuum damage mechanicsI and II, J. Appl. Mech., Trans. ASME 55 (1988) 5572.
[13] D. Krajcinovic, G.U. Fonseka, The continuous damage theory of brittle materialsI and II, J. Appl. Mech.,
Trans. ASME 48 (1981) 809.
[14] D. Krajcinovic, Constitutive equation for damaging materials, J. Appl. Mech., Trans. ASME 50 (1983) 355.
[15] D. Krajcinovic, S. Selvraj, Creep rupture of metalsan analytical model, J. Engng. Mater. Technol., Trans. ASME
106 (1984) 405.
[16] R. Ilankaman, D. Krajcinovic, A constitutive theory for progressively deteriorating brittle solids, Int. J. Solids
Struct. 23 (1987) 1521.
[17] E.T. Onat, F.A. Leckie, Representation of mechanical behavior in the presence of changing internal structures,
J. Appl. Mech., Trans. ASME 55 (1988) 1.
[18] J. Betten, Damage tensor in continuum mechanics, J. Mec. Theor. Appl. 2 (1983) 13.
[19] C.L. Chow, T.J. Lu, On evolution laws of anisotropic damage, Engng. Fract. Mech. 34 (1989) 679.
[20] J.C. Simo, J.W. Ju, Strain- and stress-based constitutive damage modelsI and II, Int. J. Solids Struct. 23 (1987)
821869.
[21] J.W. Ju, On energy-based coupled elastoplastic damage theories: constitutive modeling and computational aspects,
Int. J. Solids Struct. 35 (7) (1989) 803833.
[22] S. Murakami, Notion of continuum damage mechanics and its application to anisotropic creep damage theory,
J. Engng. Mater. Technol. 105 (1983) 99105.
[23] M. Stumvoll, G. Swoboda, Deformation behavior of ductile solids containing anisotropic damage, J. Engng. Mech.
ASCE 119 (1993) 13311352.
[24] S.C. Cowin, The relationship between the elasticity tensor and the fabric tensor, Mech. Mater. 4 (1985) 137147.
[25] G. Swoboda, Q. Yang, An energy-based damage model of geomaterialsI. Formulation and numerical results.
II. Deduction of damage evolution laws, Int. J. Solids Struct. 36 (1999) 17191755.
[26] M.F. Kanninen, C.H. Popelar, Advanced Fracture Mechanics, Oxford University Press, New York, 1985.
[27] R. Yang, W.F. Bawden, P.D. Katsabanis, A new constitutive model for blast damage, Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. 33 (1996) 245254.
[28] S. Murakami, K. Kamiya, Constitutive and damage evolution equations of elastic-brittle materials based on
irreversible thermodynamics, Int. J. Mech. Sci. 39 (1997) 473486.
[29] F.R. Tuler, B.M. Butcher, A criterion for the time dependence of dynamic fracture, Int. J. Fract. Mech. 4 (1968)
431437.
[30] B. Steverding, S.H. Lehnigk, The fracture penetration depth of stress pulses, Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. 13 (1976) 7580.
[31] D.E. Grady, M.E. Kipp, Dynamic rock fragmentation, in: B.K. Atkinson (Ed.), Fracture Mechanics of Rock,
Academic Press, London, 1987, pp. 429475.
[32] M.L. Green, Laboratory tests on salem limestone, in: Report of Geomechanical Division, Structures Laboratory,
Waterways Experiment Station (WES), Department of Army, Vicksburg, MS, 1992.
[33] R.L. Parrish, J.S. Kuszmaul, Development of a predictive capability for oil shale rubblization: results of recent
cratering experiment, in: Seventeenth Oil Shale Symposium Proceedings, Golden, CO, 1984.
[34] J.S. Kuszmaul, Numerical modeling of oil shale fragmentation experiments, in: Proceedings of the Society of
Explosive Engineers, Mini Symposium on Blasting Research, San Diego, CA, 1985.
[35] Autodyn User Manual, revision 3.0, Century Dynamics, San Ramon, CA, 1997.

You might also like