You are on page 1of 39

Journal of the Mechanics and Physics of Solids

54 (2006) 12371275
Impact initiation of explosives and propellants via
statistical crack mechanics
J.K. Dienes, Q.H. Zuo

, J.D. Kershner
Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
Received 3 May 2005; received in revised form 13 October 2005; accepted 2 December 2005
Abstract
A statistical approach has been developed for modeling the dynamic response of brittle materials
by superimposing the effects of a myriad of microcracks, including opening, shear, growth and
coalescence, taking as a starting point the well-established theory of penny-shaped cracks. This paper
discusses the general approach, but in particular an application to the sensitivity of explosives and
propellants, which often contain brittle constituents. We examine the hypothesis that the intense
heating by frictional sliding between the faces of a closed crack during unstable growth can form a
hot spot, causing localized melting, ignition, and fast burn of the reactive material adjacent to the
crack. Opening and growth of a closed crack due to the pressure of burned gases inside the crack and
interactions of adjacent cracks can lead to violent reaction, with detonation as a possible
consequence.
This approach was used to model a multiple-shock experiment by Mulford et al. [1993. Initiation
of preshocked high explosives PBX-9404, PBX-9502, PBX-9501, monitored with in-material
magnetic gauging. In: Proceedings of the 10th International Detonation Symposium, pp. 459467]
involving initiation and subsequent quenching of chemical reactions in a slab of PBX 9501 impacted
by a two-material yer plate. We examine the effects of crack orientation and temperature
dependence of viscosity of the melt on the response. Numerical results conrm our theoretical nding
[Zuo, Q.H., Dienes, J.K., 2005. On the stability of penny-shaped cracks with friction: the ve types of
brittle behavior. Int. J. Solids Struct. 42, 13091326] that crack orientation has a signicant effect on
brittle behavior, especially under compressive loading where interfacial friction plays an important
role. With a reasonable choice of crack orientation and a temperature-dependent viscosity obtained
ARTICLE IN PRESS
www.elsevier.com/locate/jmps
0022-5096/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jmps.2005.12.001

Corresponding author. Tel.: +1 505 667 0377; fax: +1 505 665 5926.
E-mail address: zuo@lanl.gov (Q.H. Zuo).
from molecular dynamics calculations, the calculated particle velocities compare well with those
measured using embedded velocity gauges.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Penny-shaped cracks; Frictional heating; Melting and ignition; Initiation and quenching of reaction;
Explosion
1. Introduction
Explosives and propellants may react violently as a result of relatively mild stimuli for
which the continuum heating is negligible (Jensen et al., 1981; Green et al., 1981; Idar et al.,
1998). For example, in a test by Idar et al. (1998), an explosive (PBX 9501) shows evidence
of chemical reactions at impact speeds below 50 m/s, and it undergoes violent reactions at
somewhat higher impact velocity. This anomalous behavior (XDT) has been of great
concern to designers and users of explosives and rockets. To quantify this sensitivity and
assess the risks associated with reactive materials, the actual mechanisms of initiation must
be known in some detail. Since the temperature rise due to mechanical dissipation during
uniform continuum deformation is too low to initiate reactions, we can conclude that the
initiation is localized in small volumes (hot spots) where the heating is intense enough to
lead to a vigorous reaction. Numerous localization processes have been proposed to
account for the formation of hot spots that initiate reactions (e.g., Field et al., 1992;
Bonnett and Butler, 1996), but for impacts at very low speeds it seems likely to us that
interfacial friction in closed cracks is the dominant mechanism. For completeness the 10
mechanisms noted by Field et al. are listed in Table 1, but crack friction is not among
them. However, crack formation in propellant and explosives is often observed, both
macroscopically and in micrographs (Howe et al., 1985; Skidmore et al., 1997). This may
occur as a result of initial formulation or subsequent damage. One occasionally reads that
fracture does not lead to initiation (Chaudri, 1972; Balzer et al., 2002), but the evidence
concerns open cracks, not closed cracks where friction plays a role. When cracks are closed
we nd that the heat generated by interfacial friction can cause a signicant reaction. The
ARTICLE IN PRESS
Table 1
The hot-spot mechanisms cited by Field et al. (1992) (in condensed form)
1. Adiabatic compression of cavity gases
2. Heating of solid adjacent to collapsing cavity
3. Viscous heating of binder between grains
4. Friction between impacting surfaces
5. Localized adiabatic shear
a
6. Heating at crack tips
7. Heating at dislocation pile ups
8. Spark discharge
9. Triboluminescent discharge
10. Decomposition followed by Joule heating of metallic elements
a
Adiabatic shear is sometimes considered to be similar to shear cracking, but there is an essential difference in
the details, with the former usually involving an instability due to thermal softening and the latter a strictly
mechanical instability involving a competition of strain energy and surface energy, with very high crack speeds
possible.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1238
interfacial friction mechanism discussed here accounts for the XDT incidents reported by
Jensen et al., Green et al., and Idar et al. Moreover, the multiple-shock experiment of
Mulford et al. (1993) that led to reactions following a second shock can be accurately
simulated with this mechanism.
Our calculations represent a feasibility study to determine whether the proposed fracture
and friction mechanisms would explain the observations, while we also considered whether
other mechanisms might not explain the observed details. Observations of particular
concern for XDT are: (a) that it occurs at very late times (hundreds of microseconds rather
than just a few for shock-to-detonation transition (SDT); (b) it can be more violent than
SDT (greater air-blast pressures); (c) large samples are more sensitive than small ones (size
effect); and (d) violent reactions occurred in only 12 out of 50 shots in those rst propellant
tests of Jensen et al. (the remainder involved mild deagrations).
All these features can be accounted for with the shear-crack heating mechanism and
crack statistics which are included in our material model (Statistical CRAck Mechanics,
SCRAM, Dienes, 1978, 1985, 1996). The model accounts for the opening, shear, growth,
and coalescence of an ensemble of penny-shaped cracks, as well as plastic ow and a
nonlinear equation of state. To characterize the response of reactive materials, the model
also accounts for heating produced by interfacial friction on shear (closed) cracks, and
possible melting, ignition and fast burning of the material next to the crack surfaces. In
addition to studying explosive sensitivity (Dienes, 1982, 1984, 1996), the model has been
used to study in situ retorting of oil shale and it was shown that the anisotropy of the
bedded rock accounted for the formation of an aspirin-shaped cavity formed by a spherical
charge (Dienes, 1981). It has also been used to model damage and failure of a ceramic
armor under ballistic impact (Meyer et al., 1999; Zuo et al., 2003).
Based on SCRAM, our colleagues at Los Alamos have developed two simplied models
for damage and failure of brittle materials. Addessio and Johnson (1990) proposed an
isotropic damage model (ISO-SCRAM) for the dynamic response of brittle materials
under nearly isotropic stress states by assuming that the crack distribution remains
isotropic during the deformation. Their calculations compared favorably with shock
compression and release experiments of plate impact for three ceramics. Based on SCRAM
and ISO-SCRAM, Bennett et al. (1998) and Hackett and Bennett (2000) explored a
mechanicalthermal model (Visco-SCRAM) for ignition of PBX 9501 under non-shock
impacts. They applied it to simulate a non-shock ignition experiment by Asay et al. (1997)
in which a small piece of conned PBX 9501 is impacted by a steel plunger and the
displacements and temperatures on the surface are measured. The computed in-plane
surface-displacement eld matches the measured eld reasonably well. Their results fully
support the notion of Dienes (1982, 1984, 1996) that frictional heating can cause ignition
and initiation in propellants and explosives under weak stimuli. Those simplied models
(ISO-SCRAM and Visco-SCRAM) assume that the crack distribution remains isotropic
during the deformation; consequently, they do not account for the anisotropic nature of
damage. SCRAM accounts for material anisotropy, either originally present in the
material (such as bedding cracks in oil shale) or induced by cracking, by tracking the
evolution of crack sizes in various directions. Several other researchers have also
considered anisotropic distribution of microcracks in their models for damage and failure
of brittle materials (e.g., Curran et al., 1993; Espinosa, 1995; Gailly and Espinosa, 2002).
Another feature of SCRAM, which is not included in ISO-SCRAM and Visco-SCRAM, is
shear dilatancy. Shear dilatancy can result from opening of cracks with certain orientations
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1239
under shear, or from joint opening due to asperities and particles, and is an important
aspect of the dynamic response of brittle materials (Scholz, 2002). Joint or crack opening
as a source of dilatancy has been incorporated into SCRAM and plays a role in the
response of ceramic armor to impact (Zuo et al., 2003).
In addition to accounting for anisotropic cracking and shear dilatancy, SCRAM
accounts for several thermalmechanicalchemical effects which are important to the
modeling of damage and initiation of reactive materials. (1) Crack coalescence is
accounted for by dividing the cracks into 2 types: active and inactive. Active (isolated)
cracks can grow under moderate stresses, while inactive (connected) cracks do not grow
because of intersections with other cracks. The evolution of active and inactive crack sets is
given by a Liouville equation which can be solved analytically when the initial distribution
is exponential (Dienes, 1985). (2) The effect of heating by frictional sliding of crack faces
on the thermal response (i.e., melting, ignition, burning, explosion) is accounted for by a
subgrid model in which the ux due to frictional heating serves as a boundary condition to
a one-dimensional heat equation with an Arrhenius source term representing chemical
reactions. The model accounts for melting of the material adjacent to a shear-crack surface
by allowing transition of the mechanical heating mechanism from solid frictional to
viscous shearing of the molten layer when the crack surface is completely melted. The
heating from viscous shearing can be substantial due to the high shear rate in the very thin
molten layer next to the crack surface. (3) The effects of latent heat of melting are
accounted for by allowing a region of mixed solid and liquid phases where the continuum
temperature remains at the melting point. The kinetics of phase transition is governed by
the ratio of volumetric heat generated and the latent heat. The latent heat affects the
spread of the melting front from the crack surface into the bulk solid, and is accounted for
in the model by solving the Stefan problem (Dienes et al., 2002). (4) The intense viscous
heating in the molten layer can bring the local temperature to the ignition point. The
burning process is represented by the burn model of Ward et al. (1998) (WSB). The mass
ux predicted by WSB is used to calculate the pressure inside the burned gases, which is a
part of the loading used in our crack dynamics calculation.
One of the main differences between SCRAM and Visco-SCRAM is that SCRAM
models the complete process of initiation in which the late-stage fast burn following
ignition plays an important role whereas Visco-SCRAM is only intended for modeling the
response leading up to ignition. As such, many of the important effects discussed above are
either not considered (e.g., burning of cracks following ignition and the effects of gas
pressure in the reaction products on crack responses), or accounted for by some rough
approximations (e.g., approximating the interfacial frictional heating on crack surfaces as
a volumetric heating source, instead of a ux boundary condition as done in the current
model and consistent with physics).
In this paper, we will show that in the multiple-shock experiment of Mulford et al.
(1993) the particle velocities calculated by SCRAM agree well with the data for several
(eleven) locations inside the PBX 9501 target where the measurements were taken. We will
also show that crack orientation and the temperature-dependent viscosity model have
important effects in the material response. We continue to seek out improvements to
the modeling of impact initiation, but selecting the most critical issue is difcult without
good experimental guidance. We will recommend an experiment involving a mixture of
inert and HMX (energetic) grains that would help clarify the mechanics and statistics of
hot spots.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1240
The paper will proceed as follows. A fairly detailed description of the SCRAM model
including both mechanics and thermalchemical analysis is given in Section 2. Two model
problems will be provided in Section 3 as a verication of our theory and numerical
implementations. In Section 4, we will apply the model to PBX 9501, showing rst a t of
the model to the stressstrain responses measured by Wiegand (1998a, b), then a
comparison of the calculated initiation results with the multiple-shock experiments of PBX
9501, using the material properties determined from Wiegands mechanical tests. In
Section 5, we will discuss the effects of sample size and defect statistics on the safety of
explosives and propellants, and recommend, for future work, a new test, intended to
examine microscopically the hot-spot mechanism used in the paper, namely, frictional
heating from sliding of shear cracks. A summary of the paper and some concluding
remarks are given in Section 6.
1.1. Notation
For compactness, the following direct notation for vector and tensor operations (e.g.
Gurtin, 1981) will be used in most of the paper:
i d
ij
e
i
Q e
j
; I
1
2
(d
ik
d
jl
d
il
d
jk
)e
i
Q e
j
Q e
k
Q e
l
,
u Q v u
i
v
j
e
i
Q e
j
; A Q B A
ij
B
kl
e
i
Q e
j
Q e
k
Q e
l
,
Au A
ik
u
k
e
i
; AB A
ik
B
kj
e
i
Q e
j
; CB C
ijkl
B
kl
e
i
Q e
j
,
u v u
k
v
k
; A : B tr(A
T
B) = A
ik
B
ik
,
A
.
B
1
2
(A
ik
B
jl
A
il
B
jk
)e
i
Q e
j
Q e
k
Q e
l
,
where i is the second-order identity; I, the fourth-order identity tensor; d
ij
, the Kronecker
delta; {e
i
](i = 1; 2; 3), an arbitrary basis; u, v, vectors; A, B, symmetric, second-order
tensors; C, fourth-order tensor.
2. Statistical crack mechanics model
It is generally known that materials can exhibit either ductile or brittle behavior, and
that the type of behavior depends on temperature, strain rate, and stress state. In SCRAM,
we develop a general theory that subsumes both kinds of behavior by superimposing the
strain rates due to various physical processes. This is a generalization of the idea of Reuss
that elastic and plastic strain rates should be superimposed, i.e.,
D =

a
D
a
, (1a)
where D represents the symmetric part of the velocity gradient (the stretching, e.g., Gurtin,
1981) and D
a
is the contribution of the physical mechanism of type a. The premise of
Eq. (1a) is that deformation is often the consequence of numerous independent physical
processes which can be superimposed to obtain an overall deformation. This is illustrated
schematically in Fig. 1. The difference in velocity between two points P
1
and P
2
, can be
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1241
written as
Du u
2
u
1
=

Du
c

Du
d
, (1b)
where the rst sum is taken over the continuous regions between the defects, and the
second sum is taken over the defects between continuous regions. The detailed derivation
of Eq. (1a) from Eq. (1b), which is carried out for arbitrarily large deformation, has been
given by Dienes (1989, 1996, 2003) and is somewhat involved. In its current form the
superposition accounts for the opening, shear, growth and coalescence of penny-shaped
cracks, and plastic ow
D = D
m
D
c
D
g
D
p
, (2)
where D
m
denotes the matrix (elastic) stretching (or the strain rate for small deformations);
D
c
, the contribution from opening and shearing of cracks (present for both stationary and
growing cracks); D
g
, the contribution due to crack growth; and D
p
, the plastic stretching
(strain rate). There is an extensive literature discussing the appropriateness of the additive
decomposition of the strain rate in the context of elasto-plasticity theory (e.g. Simo and
Hughes, 1998). We only note that for small elastic strain (compared to one), the additive
superposition of strain rates Eq. (2) is consistent with the multiplicative decomposition of
the deformation gradient (e.g. ABAQUS, 1998). For brittle materials considered in this
paper the elastic strains are always much smaller than one. Since the stress power is
_ e = r : D, each term of Eq. (2) is associated with the energetics of a particular deformation
mechanism. In a more general realization of the SCRAM algorithm, spherical voids,
anisotropic high-pressure effects, and twinning, for example, could be added to the current
formulation.
2.1. Crack strain rate, D
c
, and added compliance
Consider an ensemble of penny-shaped microcracks randomly distributed within a
statistically homogeneous volume of a brittle material under multiaxial loading. The
requirements for a material volume to be statistically homogeneous are those discussed by
ARTICLE IN PRESS
Fig. 1. A conceptual diagram of the defects assumed in deriving the generalized superposition of strain rates
described by Eq. (1a). Detailed derivation is given by Dienes (1989, 1996, 2003).
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1242
Krajcinovic (1998). The contribution to the stretching of a crack set, i.e., a homogeneous
distribution of similar penny-shaped cracks with the unit normal n and radius c, is (Dienes,
1989, 1996)
DD
c
(c; n) = bn(c; n; t)c
3
(d
n
d
t
)DcDC, (3a)
d
n
= (2 n)H[s
n
](n ^ rn)n Q n, (3b)
d
t
= 1 Z)(( ^ rn) Q n n Q ( ^ rn) 2(n ^ rn)n Q n), (3c)
where r is the far-eld (applied) Cauchy stress, and ^ r represents the GreenNaghdi (polar)
rate i.e., ^ r _ r rX Xr, with X representing the rate of material rotation (e.g., Dienes,
1979, 1996; Belytschko et al., 2000). The elastic constant b 8(1 n)=[3G(2 n)], arises
from analytic solutions for open and closed penny-shaped cracks (Sack, 1946; Segedin,
1950; Keer, 1966), with G and n the elastic shear modulus and Poissons ratio of the
isotropic matrix (undamaged) material, respectively. n(c; n; t) is the number density
dening the distribution of crack radii and orientations, which evolves with the time t.
That is, n(c; n; t)DcDC represents the number density of cracks (number of cracks per unit
volume) whose radii are between c and c Dc, and have a unit normal within a small solid
angle DC around n (e.g., Oda, 1983; Dienes, 1985).
In Eq. (3b), s
n
n rn is the normal component of the remote (far-eld) traction and H
is the Heaviside function (one for positive arguments and zero for negative arguments) so
that H[s
n
] = 1 for an open crack and H[s
n
] = 0 for a closed crack. The quantity Z in Eq.
(3c), arising from the effects of interfacial friction between closed-crack faces (reducing the
interfacial sliding), is given by
Z
ms
n
)
s
n
, (3d)
where s
n
[n r
2
n (n rn)
2
]
1=2
is the shear component of the remote traction and m the
friction coefcient (static or dynamic depending on whether the crack is sliding or at rest).
The angled brackets in Eqs. (3c) and (3d) denote the Macaulay bracket, which takes the
value of the argument when positive and is zero otherwise. The Macaulay bracket in
Eq. (3c) is used so that sliding of the crack faces (hence stretching due to crack shearing) is
prohibited when the crack is locked by friction ( ms
n
4s
n
, or Z41). This introduces the
hysteretic effect of solid friction. It follows from the denition that Z = 0 for an open crack
(s
n
40), and 0p1 Z)p1 for a closed crack. For certain orientations, closed cracks may
be friction-locked, i.e., 1 Z) = 0; then both d
n
and d
t
vanish.
For closed (shear) cracks (s
n
p0), it follows from Eqs. (3a)(3c) that
DD
c
(c; n) = bn(c; n; t)c
3
DcDCd
t
= 1 Z)bn(c; n; t)c
3
DcDCc
s
^ r. (4a)
The shear fabric tensor (fourth-order) c
s
is (Oda et al., 1984)
c
s
b 2a, (4b)
where
a n Q n Q n Q n, (4c)
b i
.
(n Q n) (n Q n)
.
i. (4d)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1243
In Eq. (4d), the operator
.
is the cross-composition of two symmetric second-order
tensors dened in Notation.
Similarly, for open cracks (s
n
40), Eqs. (3a)(3c) reduce to
DD
c
(c; n) = bn(c; n; t)c
3
DcDCc
o
^ r (5a)
with the open fabric tensor
c
o
= (2 n)a c
s
= b na. (5b)
The total stretching due to an ensemble of microcracks of all sizes and orientations (the
crack strain rate) is obtained by summing the contributions in Eq. (3a) over all crack sizes
(0pcpo) and orientations (Dienes, 1989, 1996),
D
c
=

C;c
DD
c
(c; n) = b

C;c
n(c; n; t)c
3
(H[s
n
]c
o
(1 H[s
n
])1 Z)c
s
)DcDC^ r. (6)
In terms of the usual polar coordinates y and f, the incremental solid angle is the area of
an element on the unit sphere, DC = sin fDyDf. Cracks have symmetry such that a
reversal of 1801 leaves them unchanged. Thus, half the unit sphere is sufcient to
characterize crack orientation; consequently, the integration limits for orientation are
0pyp2p and 0pfpp=2.
2.1.1. Added compliance
It follows from Eq. (6) that the stretching due to an ensemble of cracks is the product of
the added compliance and the stress rate:
D
c
= (C
o
C
s
) ^ r. (7)
The added compliance due to open and closed (shear) cracks can be written as
C
o
= b

C
H[s
n
]F(n; t)c
o
DC, (8a)
C
s
= b

C
(1 H[s
n
])1 Z)F(n; t)c
s
DC, (8b)
where the summation is taken over crack orientations. In SCRAM, the continuous
distribution of crack orientations n is approximated with a number of discrete orientations
(crack bins). Crack orientations range over 2p, half a unit sphere, and one particularly
convenient discretization scheme is to divide the hemisphere into N elements with equal
areas so that DC = 2p=N.
In Eqs. (8a) and (8b), the function F(n; t) is the third moment of the density-distribution
function n(c; n; t):
F(n; t)
_
o
0
n(c; n; t)c
3
dc. (9)
This function comes out of the analysis (Dienes, 1989) and was termed effective crack
volume previously (Dienes, 1996), though it is not the real volume of a crack, which is
zero for cracks with zero thickness. Note that the sum over crack radii given by Eq. (6) has
been replaced by integrals over a continuous distribution of crack radii. Analytic
expressions for the integrals are developed in Section 2.4.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1244
Each crack orientation (bin) n
a
(a = 1; . . . ; N) has its own number density distribution
function, n(c; n
a
; t), and anisotropic damage is captured by tracking the evolution of those
distribution functions with time. Cracks within the crack bin a are either open or closed
depending on the sign of s
n
a = n
a
rn
a
for the bin; the third moment F(n; t) in Eq. (9) for
the crack bin contributes either to C
o
[Eq. (8a)] or C
s
[Eq. (8b)], accordingly. In general,
whether a crack bin is open or closed depends on the stress state and the crack orientation
of the bin, and some bins are open and some are closed. Under certain stress states,
however, the status (open or closed) of a crack bin is independent of the bin orientation.
For example, when the principal stresses are all positive (tensile), all crack bins are open.
Consequently,
C
o
= b
2p
N

a=1;N
F(n
a
; t)c
o
(n
a
), (10a)
C
s
= o. (10b)
Conversely, when the principal stresses are all negative (compressive), all crack bins are
closed, and
C
o
= o, (11a)
C
s
= b
2p
N

a=1;N
1 Z(n
a
))F(n
a
; t)c
s
(n
a
). (11b)
Furthermore, for certain stress states where the principal stresses are all compressive and
close to each other, cracks in all bins are closed and locked by friction; consequently, there
is no added compliance (or damage) due to cracks. This dependency of the material
compliance on the stress state does not occur in a truly linear material model and implies
that the material model is nonlinear, even though linear fracture mechanics is used to
derive the crack opening and sliding. The range of the stress states for which all cracks are
friction-locked increases with the friction coefcient and has been given by Zuo and Dienes
(2005) and Dienes et al. (2004).
2.2. Stretching due to crack growth, D
g
Cracks can contribute to the stretching by growing in size. The stretching due to the rate
of growth of open cracks, _ c, is
D
o
g
=
_
C
o
r, (12a)
_
C
o
= b

C
H[s
n
]
_
F(n; t)c
o
DC. (12b)
Similarly, for closed cracks,
D
s
g
=
_
C
s
r, (13a)
_
C
s
= b

C
(1 H[s
n
])1 Z)
_
F(n; t)c
s
DC. (13b)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1245
It follows from Eq. (9) that the rate of the third moment F(n; t) is given by
_
F(n; t) =
_
o
0
qn(c; n; t)
qt
c
3
dc. (14)
The relationship between
_
F(n; t) and _ c is given in Section 2.4. The stretching due to the
growth of open and closed cracks is then
D
g
= D
o
g
D
s
g
= (
_
C
o

_
C
s
)r. (15)
The rate of change of the added compliance (or material damage) is the sum of
contributions of open and closed cracks
_
C
o

_
C
s
= b

C
_
F(n; t)(H[s
n
]c
o
(1 H[s
n
])1 Z)c
s
)DC. (16)
It is interesting to observe that the stretching due to crack growth is like a viscous strain
rate in that it is proportional to the stress tensor, not its rate. Crack growth and
coalescence are discussed next.
2.3. Crack growth
The classic theory of crack instability and growth assumes that cracks grow at high
speed when the applied stress exceeds a critical level causing cracks to become unstable
(Freund, 1990). Practical observations on a variety of materials reported by Stroh (1957)
show that crack tip speeds can approach roughly a third the longitudinal wave speed, while
Freund (1990) argues on theoretical grounds that they approach the Rayleigh wave speed
at high stress, somewhat faster than the Stroh result. On the other hand, it has been
observed (Charles, 1958; Evans, 1974; Kanninen and Popelar, 1985) that, when the stress is
below the critical level for instability, cracks may grow at low speeds. (This regime may be
governed by thermally activated diffusion rather than dynamic processes.) The empirical
results can be expressed as
_ c = c
R
g(r; n; c)
g
1
_ _
n
for g(r; n; c)pg
tr
, (17a)
where g(r; n; c) is the energy-release rate for the crack with normal n and size c. The
terminal crack speed c
R
is the Rayleigh wave speed, and n is a model parameter typically
between 5 and 10. The constants g
1
and g
tr
are related to the effective surface energy g and
the parameter n (Dienes and Kershner, 1998), and are given next.
The above result applies when the stress is below a critical value. At high stress, when
cracks are unstable, Freund (1990) gives the result
_ c = c
R
1
g
c
g(r; n; c)
_ _
for g(r; n; c)Xg
c
, (17b)
where g
c
= 2 g is the critical energy-release rate, twice the effective surface energy g. It
follows that for an unstable crack (g(r; n; c)Xg
c
), the crack growth speed asympto-
tically approaches the terminal crack speed c
R
as the energy release rate increases. The
parameter g
tr
controls where the slow, diffusion-controlled crack growth transitions to
fast, dynamic crack growth. g
1
and g
tr
are found by requiring that the magnitude and
slope of the curves dened by Eqs. (17a) and (17b) be continuous at the transition point
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1246
where g(r; n; c) = g
tr
g
tr
= 1
1
n
_ _
g
c
4g
c
, (18a)
g
1
= (n 1)
1=n
1
1
n
_ _
g
c
4g
tr
. (18b)
For a typical value of n = 6, the transition occurs at g
tr
= (7=6)g
c
, slightly above the
critical energy-release rate, that is, the slow crack growth formulation extends into the
unstable crack regime by a small amount. At the cross-over, g(r; n; c) = g
tr
, the crack
growth speed is _ c = c
R
=(n 1), which is (1=7)c
R
for n = 6. In this approach, we no longer
deal with crack instability, but with a transition from slow to fast crack growth. However,
the growth is very slow for stable cracks (g(r; n; c)og
c
). It follows from Eq. (17a) that for a
stable crack bin,
_ co
g
c
g
1
_ _
n
c
R
=
n
n
(n 1)
n1
c
R
. (19)
For n = 6, the crack growth speed is _ co0:057c
R
for stable cracks.
The current formulation is a generalization of Dienes and Kershner (1998) in which the
stress intensity factor is used for calculating crack growth, rather than the energy-release
rate used here. When the crack is under mixed-mode loading (mode-I, and both mode-II
and mode-III for a penny-shaped crack), it is more convenient to use the current
formulation based on the energy-release rate. In the work of Dienes and Kershner (1998)
and Bennett et al. (1998), since only closed cracks are considered, the two formulations are
equivalent.
For a penny-shaped crack, the energy-release rate can be written as (Rice, 1984)
g(r; n; c) =
4
p
(1 n)
(2 n)
f (r; n)c
G
. (20)
The expression for the stress function f (r; n) depends on whether the crack is open (the
normal component of traction is tensile) or closed (the normal component is compressive
and controls the interfacial friction). For an open crack (s
n
40), both normal and shear
stresses contribute to crack instability and the stress function f (r; n) is (Keer, 1966)
f (r; n) = 1
n
2
_ _
s
2
n
s
2
n
, (21a)
where s
n
and s
n
are the normal and shear components of the remote traction dened in
connection with Eq. (3d).
For a closed crack (s
n
p0), the friction on the crack surface is stabilizing. If the
Coulomb friction law is assumed, then the stress function f (r; n) is (Rice, 1984)
f (r; n) = s
n
ms
n
)
2
. (21b)
When the crack is friction-locked (s
n
o ms
n
), it follows from Eqs. (20) and (21b) that
f (r; n) = 0 and g(r; n; c) = 0. Thus, the crack is friction-locked and remains stable and
stationary (_ c = 0) for any applied stress.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1247
2.4. Crack coalescence
If cracks grew without coalescence the statistical problem would be trivial but, in fact,
crack intersections are a vital part of material failure. Without coalescence the largest
crack would be the most unstable, and it would grow without bound, separating the
material rapidly into 2 parts while the others are unaffected. This may be observed in some
circumstances, but SCRAM is concerned with problems where damage occurs as a result
of large numbers of microcracks. Failure in granular materials such as explosives,
propellants, and ceramics is the result of microcrack growth and coalescence, and no single
microcrack dominates. This is consistent with the observation that the stressstrain curves
for brittle materials often show a gradual increase of compliance as the stress increases.
Furthermore, we observe crack networks with complex structure in micrographs of
damaged materials. This growth and coalescence of cracks result, ultimately, in
fragmentation if the loading is of sufcient duration and intensity, notably in spall
(Antoun et al., 2003).
The rst stage in formulating SCRAM crack statistics is to divide cracks into two
categories, active (isolated) and inactive (connected). Active cracks are capable of growth
as a result of either the instability that occurs at stresses above the critical value (fast
growth) or the thermally activated processes (slow growth). Inactive cracks are not capable
of growth because they have intercepted a signicant number (usually 34) of other cracks,
thereby eliminating large portions of the regions of high stress at the crack edge. These
inactive cracks are the ones we consider to form the faces of fragments. Of course, in real
fragments the faces are polygons rather than circles, but we take the circles as a rough
approximation to the true area. Active and inactive cracks can be observed in most
geological formations-active cracks are like the preceding dash and inactive cracks are
observed as something like a T or H on the face of an outcropping or a rock (Dienes,
2005). Active cracks have a statistical distribution L(c; n; t) representing the average
number of cracks per unit volume exceeding c in radius, while inactive cracks have a
distribution M(c; n; t). The dependence on time, t, governs only during periods of crack
growth.
In terms of the distribution functions, L(c; n; t) and M(c; n; t) are
L(c; n; t) =
_
o
c
n
act
(c; n; t) dc, (22a)
M(c; n; t) =
_
o
c
n
ina
(c; n; t) dc. (22b)
Since a crack is taken to be either active or inactive, the total number of cracks is the sum
of the cracks in the active and inactive sets:
n
act
(c; n; t) n
ina
(c; n; t) = n(c; n; t). (23c)
It is shown by Dienes (1978, 1985) that the distributions are related by a Liouville equation
qL
qt
_ c
qL
qc
=
qM
qt
, (24)
which holds for each crack orientation (n) and does not involve other crack orientations
because cracks in various orientations are assumed to remain plane and to grow
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1248
independently. For compactness, the explicit dependence on crack orientation and time is
dropped in this section when the context is appropriate. In the following, the crack speed _ c
is assumed constant in time and crack size (it still varies with the crack orientation), though
it is straightforward to handle the case when it is variable (Dienes, 1985).
Based on experimental observations (Seaman et al., 1976; Curran et al., 1987; Curran
and Seaman, 1996; Scholz, 2002), it is assumed here that the initial distribution is
exponential in crack size (Dienes, 1985)
L(c; n; 0) = L
0
(n) exp(c= c
0
(n)), (25)
where c
0
(n) and L
0
(n) are the mean initial crack size (radius) and total number of cracks
per unit volume per 2p, respectively. It is typically assumed that the crack number density
and mean crack size are isotropic, initially, though the theory allows for an initially
anisotropic distribution of cracks. For oil shale an additional set of large cracks in the
bedding planes was generated (Dienes, 1981).
If the crack density is not too large, the rate at which cracks coalesce is proportional to
L(c; n; t),
qM
qt
= kL, (26a)
where k is a constant that depends on the initial distribution of cracks. An analysis of the
mean free path given by Dienes (1989) leads to the expression for the coalescence constant:
k = 2p
4
L
0
c
2
0
_ c= a. (26b)
The constant a is the number of intersections required to convert an active crack to one
that is inactive (termination of crack growth). It can be estimated on intuitive grounds to
lie between 3 and 4 [analysis of one series of experiments leads to 3.4 (Dienes, 2005)]. Since,
in general, L
0
, c
0
, _ c depend on the crack orientation n, the constant k is in general
orientation dependent, k = k(n), though in practice it has been assumed constant.
With the initial distribution of cracks given by Eq. (25), the Liouville equation has the
analytical solution:
L = 0; M =
1

b
[(_ c= c
0
)e
kc=_ c
ke
c= c
0


be
kt
]L
0
for co_ ct, (27a)
L = L
0
e
(c_ ct)= c
0
kt
; M =
k

b
(e

bt
1)e
c= c
0
L
0
for c4_ ct (27b)
with

b _ c= c
0
k. The total number of cracks should, and does, remain equal to L
0
. To
see this note that the number of active cracks with radius exceeding zero is the same as the
number exceeding _ ct, L = L
0
e
kt
. The number of inactive cracks reduces, for c = 0, to
M = L
0
(1 e
kt
), so that the sum is indeed L
0
. Recall that the time t here is the duration
of the crack growth, which can be signicantly less than the total elapsed time.
Let F
L
(n; t) and F
M
(n; t) be the contributions from active and inactive cracks to the third
moment
F
L
(n; t) =
_
o
0
L
c
(n; c; t)c
3
dc, (28a)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1249
F
M
(n; t) =
_
o
0
M
c
(n; c; t)c
3
dc, (28b)
where the subscript c denotes the derivative with respect to crack size. Analytical
expressions for F
L
(n; t) and F
M
(n; t) have been found: for active cracks,
F
L
(n; t) = L
0
exp(kt)(^ c
3
3 c
0
^ c
2
6 c
2
0
^ c 6 c
3
0
), (29a)
where
^ c _ ct = c(n; t) c
0
(29b)
is the extent of crack growth for the crack orientation n, and c(n; t) is the current mean
crack size for each orientation. For inactive cracks,
F
M
(n; t) =
L
0
c
0

b
[6(~ c
4
c
4
0
) exp(kt)H], (30a)
where ~ c _ c=k and
H ^ c
3
( c
0
~ c) 3^ c
2
( c
2
0
~ c
2
) 6^ c( c
3
0
~ c
3
) 6( c
4
0
~ c
4
). (30b)
The corresponding rates of change of the third moments are
_
F
L
(n; t) = L
0
exp(kt)[(3^ c
2
6 c
0
^ c 6 c
2
0
)_ c k(^ c
3
3 c
0
^ c
2
6 c
2
0
^ c 6 c
3
0
)], (31a)
_
F
M
(n; t) =
L
0
k
c
0

b
exp(kt)[kH (3^ c
2
( c
0
~ c) 6^ c( c
2
0
~ c
2
) 6( c
3
0
~ c
3
))_ c]. (31b)
The sums of the contributions from active and inactive cracks are the third moments
needed for calculating the added compliance and the stretching due to crack growth
F(n; t) = F
L
(n; t) F
M
(n; t), (32a)
_
F(n; t) =
_
F
L
(n; t)
_
F
M
(n; t). (32b)
When the number density of cracks becomes large it can no longer be assumed that the
rate of coalescence is constant and equal to the initial value. In that case it is still possible
to make progress toward an analytical solution, but the analysis is much more complex, as
reported by Dienes (1985, 1989). The general solution is not used in the current SCRAM
program, though the asymptotic solution is; the number of inactive cracks at late times is
used in the computer algorithm to limit crack growth.
For crack bin a (a = 1; . . . ; N), the crack growth is computed using Eq. (17) with n = n
a
and c = c(n
a
; t), the mean crack radius for the bin. The growth rate of the mean crack
radius is
_
c(n
a
; t) = _ c(r; n
a
; c(n
a
; t)); a = 1; . . . ; N, (33)
where _ c(r; n
a
; c(n
a
; t)) is given by either Eq. (17a) or Eq. (17b), depending on the energy-
release rate, g. Since the energy-release rate depends on the mean crack radius and
orientation of the bin, the crack growth rates are different for different bins, resulting in
anisotropic damage. Similarly, Eqs. (32a) and (32b) for the third moments apply to each
crack bin a with constants k,

b, and ^ c, ~ c for the bin.
It is assumed that for each crack bin, the growth rate for cracks of all sizes is the same as
that for the crack with mean crack radius of the bin, _ c =
_
c(n
a
; t). This approximation is
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1250
justied by noting that the crack distribution affects the overall behavior only through the
compliance, which involves integrals of the type (Eq. (28))

_
o
0
L
c
(c; n; t)c
3
dc. (34)
Since L
c
(c; n; t) is nearly exponential in crack size c, with mean c(n; t), the integrand has a
maximum for c = 3 c(n; t). Thus, it is only important, to rst order, to represent the
integrand correctly near c = 3 c(n; t). In particular, the distribution for co c(n; t) does not
contribute signicantly to the overall behavior.
2.5. Stress rate-stretching relationship
The matrix stretching is related to the stress rate by
D
m
= C
m
^ r, (35)
where C
m
is the compliance of the matrix material. For an isotropic, linear elastic matrix
material, C
m
= (1=(3K) 1=(2G))i Q i=3 1=(2G)I where K 2G(1 n)=(3(1 2n)) is the
bulk modulus, and I is the fourth-order identity tensor dened in Notation. Substituting
Eqs. (7) and (35) into Eq. (2) gives
D = (C
m
C
o
C
s
) ^ r D
g
D
p
. (36)
It follows that the elastic stretching is
D
e
= D
m
D
o
D
s
= C^ r, (37)
where C is the overall compliance of the material containing an ensemble of cracks:
C = C
m
C
o
C
s
. (38)
The inelastic stretching is the sum of contributions from crack growth, D
g
, and
plastic ow D
p
. In a computer program, the total stretching D and rate of material
rotation X are obtained from the momentum equation and one needs to nd the stress
rate. To that end, the constitutive law of Eq. (36) can be inverted to give the polar rate of
stress as
^ r = C
1
D
e
= C
1
(D D
g
D
p
). (39)
It is assumed here that the plastic response of the material can be represented by
kinematic hardening. Furthermore, it is assumed that the yield surface is of the form
f(r; a)
1
2
(r
d
a) : (r
d
a) Y
2
= 0, (40)
where r
d
r (tr r)i=3 is the stress deviator; a, the back stress (center of the yield
surface); :, the scalar product dened earlier; and Y, the ow stress in simple shear.
Consideration of hardening and hysteresis suggests that the center of the yield surface in
stress space evolves linearly with the plastic stretching (Prager, 1955),
^ a = bD
p
(41)
with b the kinematic hardening modulus. The plastic stretching is assumed to have
the form
D
p
=
_
l(r
d
a), (42)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1251
where
_
l is the plastic multiplier, which can be found by requiring that the stress state
remain on the yield surface during plastic deformation (the consistency condition),
^
f(r; a) = 0:
_
l =
~ r : C
1
(D D
g
)
2bY
2
~ r : C
1
~ r
, (43)
where ~ r r
d
a. With the plastic multiplier found, the plastic stretching is then obtained
by substituting Eq. (43) into Eq. (42). With the total stretching D prescribed, the plastic
stretching D
p
can be determined by using the total compliance C and the stretching due to
crack growth D
g
given by Eqs. (38) and (15), respectively.
Finally, the stress rate is
_ r = C
1
(D D
g
D
p
) Xr rX. (44)
Eq. (44) can be used to update the stress. A simple kinematic hardening model is used here,
but it is straightforward to incorporate other plasticity models when the experimental data
warrant them.
2.6. Frictional crack heating and melting
Consider the effects of heating due to crack friction on the thermo-chemical response of
a reactive material. In SCRAM, we model this heating by means of a subgrid calculation in
which the frictional heating involves only a few microns normal to the crack surface. It is
assumed that the heated zone is thin compared to the crack radius, so that the critical heat
conduction can be taken to be one dimensional and normal to the crack plane. The heating
per unit area (ux) due to interfacial friction is
_ q = ms
n
)v, (45)
where m is the solid friction coefcient dened earlier, and v is the interfacial sliding
velocity. This heating is taken as a boundary condition for the modied FrankKame-
netzky equation for one-dimensional heat conduction (Frank-Kamenetzky, 1942; Mader,
1979)
rC
v
_
T = kT
xx
r
_
Q, (46)
where T denotes the absolute temperature, r mass density, C
v
heat capacity at constant
volume, k thermal conductivity. The coordinate axis (x) is along the crack normal with the
origin at the center of the crack. The rate of heating per unit mass,
_
Q, for a reactive
material, is the sum of chemical and mechanical sources:
_
Q = Q
r
Z exp(E
A
=RT)
1
r
((1 f )Y f m_ e)_ e. (47)
The rst term on the right accounts for chemical reactions and the second for mechanical
dissipation, which is the sum of plastic work in the solid phase and viscous heating of the
melted liquid. Q
r
is the heat of reaction per unit mass for chemical decomposition;
Zexp(E
A
=RT) Arrhenius reaction rate; Z is the pre-exponential (frequency) factor; E
A
activation energy; R the universal gas constant; f (0pf p1), fraction of the phase transition
(solid to liquid) completed (f = 0 below the melting point and f = 1 for complete melting);
Y ow stress of the solid phase as before; m the viscosity of the melt; _ e the shear strain rate
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1252
in the melting or molten material adjacent to the crack surface. The thermal boundary
condition at the crack surface (x = 0) is
kT
x
= _ q. (48)
When the material adjacent to the crack surface is melting, the interface sliding velocity v
of Eq. (45) used to compute _ q is not appropriate, but, rather, the sliding velocity should be
a fraction of the crack-sliding velocity v
c
, which decreases as the fraction increases,
v = (1 f )v
c
. (49)
Since we do not know the exact relationship between the total crack sliding velocity (v
c
)
and that associated with the solid phase, a simple linear interpolation is used here. The
remainder of the crack velocity is associated with local shearing in the melting and molten
material adjacent to the crack. Thus,
_ e = fv
c
=l, (50)
where l is the width of the region at or above the melting point. (A typo in Dienes, 1996
says below rather than above.) The crack sliding velocity v
c
is computed as the central
velocity of a penny-shaped shear (closed) crack:
v
c
=
3b
2p
[_ s
n
m_ s
n
)c s
n
ms
n
)_ c], (51)
where the elastic constant b has been dened in Eq. (3a), and c is the crack radius. The rst
term is due to the change of traction on the crack surface, and the second due to crack
growth (_ c40). The second term normally dominates when the crack is unstable.
It is known that the melting point plays an important role in the sensitivity of explosives
(Bowden and Yoffe, 1952). At low temperatures (ToT
m
) the crack faces are solid, but
when the temperature at the crack interface reaches the melting point, a phase transition is
initiated. During the melting process (0of o1), the temperature remains at the melting
point (T = T
m
), and a part of the heat generated from chemical reaction and mechanical
dissipation is consumed in the phase transition. The rate of melting can be calculated from
the rate of heat generation:
_
f =
_
Q=L, (52)
where L is the latent heat of melting, and
_
Q has been given by Eq. (47). The velocity of the
melting front moving into the solid phase,
_
l(t), is given by the Stefan condition (Crank,
1984)
rL
_
l(t) = D(kT
x
), (53)
with D(kT
x
) the discontinuity in the heat ux across the melting front.
Evidence for melting in shear cracks is scant, but it has been observed in pseudotachyltes
(Scholz, 2002) and in shocked materials (Grady, 1988; Schmitt et al., 1989) and in an
impacted TNT (Howe et al., 1985). It is appropriate to mention here that there is a
distinction between shear cracks and shear bands as traditionally idealized, but this
distinction may disappear as it becomes recognized that these are two extreme cases of a
general kind of localization that involves plastic ow at one extreme and brittle failure at
the other. Shear bands in reactive materials are traditionally modeled with a one-
dimensional nonlinear equation representing softening behavior at high temperature
(Dienes, 1986), while shear cracks are typically represented with three-dimensional linear
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1253
elasticity with interfacial friction (e.g., Rice, 1984, 2000; Zuo and Dienes, 2005). This
friction is uniform over most of the crack, but nonlinear in the sense that it changes
discontinuously when the sign of the local velocity changes. The ultimate effect of the
resulting behavior, which can be a stickslip instability, can be quite complex (e.g., Heaton,
1990; Rice, 2000).
2.7. Ignition and burning of reactive materials
When melting is completed (f = 1), the heat ux ( _ q) from solid friction and the
volumetric plastic dissipation vanish. At this point the only mechanical dissipation
mechanism is viscous shearing. Normally viscous heating (m_
2
in Eq. (47)) is small for
liquids, but it is signicant in the molten layer next to a shear-crack surface because the
strain rate is very high (the thickness of the layer l is on the order of a micron in the
examples studied), especially when the crack is unstable (high v
c
). Our previous analyses of
a solid propellant showed that the cracks had grown to centimeters in length at ignition
time (Dienes, 1996). This intense viscous heating causes the temperature in the thin molten
layer to rise rapidly. Ignition occurs when the peak temperature reaches a critical value T
c
above which the Arrhenius source term in Eq. (47) becomes important, causing a rapid rise
of temperature. According to a theory of Linan and Williams (1971), the critical
temperature (ignition point) for a reactive material under constant energy ux _ q is
T
c
(T
i
; _ q) =
E
A
R
1
ln(rQ
r
ZkT
i
= _ q
2
)
, (54)
where T
i
is the initial temperature, and the other variables have already been dened.
Though the energy ux due to frictional heating normally changes with time, the critical
temperature given above does provide a useful check on our numerical scheme. The critical
temperatures for ve explosives (HMX, TNT, PBX 9501, PBX 9502, PBX 9504) for
_ q = 3060 cal=cm
2
=s and T
i
= 300 K have been calculated by Dienes (1995). The results
range from 755 K for HMX to 1179 K for TNT.
Following ignition, there is a rapid (sub-nanosecond) transition to burning, so that burn
can be considered to begin immediately when the ignition point is reached (T4T
c
). This
process can be represented by the burn model (WSB) of Ward et al. (1998). This is a two-
step model with high activation energy in the condensed phase followed by a gas-phase
reaction with vanishing small activation energy (E
g
=RT51; E
g
the activation energy of the
gas phase). Compared with the burn models assuming high activation energy for the gas
phase, the WSB model provides a better match of both the temperature prole and the
burn rate behavior for several explosives measured experimentally (Ward et al., 1998).
For our application, WSB can be briey summarized as a set of three coupled equations
for mass ux (m), surface temperature (T
s
) of the condensed phase, and the ame thickness
(x
g
). Specically:
m(T
s
) =
A
c
T
2
s
e
E
c
=T
s
E
c
(T
s
T
0
Q
c
=2)
_ _
1=2
, (55)
T
s
(m; x
g
) = T
0
Q
c

Q
g
x
g
m 1
; x
g
(m) =
2

m
2
4D
g
_
m
, (56a,b)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1254
which are Eqs. (13)(15) of WSB. In the above equations, we have followed the notation of
WSB, in which all the variables are dimensionless, and the subscripts c and g denote the
condensed and gas phases, respectively. A
c
is the reaction rate prefactor (the frequency
factor); E
c
, the activation energy; Q
c
, the chemical heat release; and T
0
, the initial
temperature for the condensed phase. Q
g
is the chemical heat release in the gas-phase
reaction. The parameter D
g
is the Damkohler number for the gas phase,
D
g
(p
g
) =
k
g
B
g
p
2
g
MW
2
c
p
(m
r
R)
2
(57)
with k
g
the thermal conductivity; B
g
, the reaction rate prefactor; p
g
, the gas pressure inside
the crack; MW, the molecular weight; c
p
, the specic heat at constant pressure; and m
r
, a
reference mass ux for the gas phase. The nonlinear coupled equations are solved by a
generalized Newtons method (Dienes and Kershner, 2001). The model constants provided
by WSB are used in our calculations, without modifying the units (MKS). The initial
pressure at ignition is given by the ideal gas law
p
i
= r
g0
RT
f
=MW, (58)
where r
g0
is the initial density of the reactive product (gas); and T
f
, the ame temperature.
The subsequent pressure is calculated assuming that the burn products following ignition
obey an adiabatic gas law
p
g
= p
i
(r
g
=r
g0
)
g
(59)
with r
g
the running (current) density of the burned gas and g c
p
=c
v
, the ratio of the
specic heats at constant pressure and volume. Consider a mass of reacted material, which
is assumed to ll a cylinder of crack radius c and height of crack roughness d
0
at the time of
ignition. The initial amount of mass per unit crack area at ignition is then
m
g0
= r
g0
d
0
. (60)
As discussed by Dienes and Kershner (2001), the initial gas density at ignition, r
g0
, is
difcult to estimate accurately due to the complexity of the process involved during high-
rate shearing and liquefaction of irregularities of crack interfaces (in geophysics such
sliding interfaces are typically separated by a layer of ne particles, Sammis and Biegel,
1989). In the current calculations, r
g0
is taken as a fraction (1/10) of the solid density and
the crack roughness is assumed to be 1 mm. During the burn, the pressure in the burned
gases opens the crack forming an oblate ellipsoid whose volume on one side of the crack is
2pc
2
(d r
b
)=3 with d the current opening of the crack and r
b
, the distance from the crack
surface to the gas-condensed phase interface. The current density is then (Dienes and
Kershner, 2001)
r
g
=
m
g
2
3
(d r
b
)
. (61)
The amount of the burned gas per unit crack area inside the cavity and the distance r
b
of
the interface during the burning are given by the mass ux
m
g
= m
g0

_
t
t
c
_
M dt, (62a)
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1255
r
b
=
_
t
t
c
_
M
r
g
dt, (62b)
where t
c
is the time of ignition, and
_
M mm
r
is the dimensional mass ux in the notation
of WSB. The gas pressure p
g
given by Eq. (59) is then used as a part of the loading in crack
dynamics calculations (Dienes, 2001; Dienes and Kershner, 2001),
r
eff
= r p
g
n Q n, (63)
where r
eff
is the effective stress, r and n, as before, the externally applied stress and the
normal of the crack, respectively.
This process continues till a critical condition when there is a transition to rapid burn, as
described by Dienes and Kershner (2001). The entire scenario is summarized in Fig. 2,
from initial shock to violent reaction.
3. Verication and validation
The SCRAM theory discussed above has been implemented into three explicit, three-
dimensional nite-element codes: PRONTO (Taylor and Flanagan, 1987), DYNA
(Hallquist and Whirley, 1989), and EPIC (Johnson et al., 2001). The details of
implementing a new constitutive model into the codes can be found in the users manuals
ARTICLE IN PRESS
Fig. 2. An overview of the stages involved in initiation according to SCRAM. (a) Shear cracks form under shock;
(b) cracks grow easily inside HMX crystals, but are inhibited by binder; (c) cracks grind and generate frictional
heat; (d) HMX reaches ignition point (T4T
c
); (e) burned gases open cracks creating high-pressure zone. Opening
is inhibited by inertia, stiffness, and damping; (f) burning increases general pressure, accelerating burn in
adjoining hot spots and (g) cracks coalesce when percolation threshold is exceeded, causing general explosion.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1256
provided by the code developers. We have paid particular attention to verifying and
validating the programming and physical concepts that are combined in SCRAM. This is
considered crucial at this point because the introduction of new physics such as the set-size
theory for percolation processes and a random crack size generator makes the damage
behavior very complex, and thus it is difcult to conrm the calculations. This is especially
the case where three-dimensional behavior is involved, and three-dimensions are required
to represent impacts that involve radial cracking and extensive damage. Radial cracking is
not a deterministic process; the cracks form at roughly random angles and may number
from 3 to more than 20. As a part of the initial condition for the material, a random
number generator is used to generate the number of cracks in each cell with each
orientation, but the statistics of crack size for each cell and orientation is based on the
exponential distribution which is discussed by Korvin (1989) and for which new evidence is
given by Antoun et al. (2003) and Scholz (2002). With this random cracking strongly
affecting the details of impact response, it is difcult to completely verify the coding, and
this is the primary reason for the needed emphasis on validation at this time.
One kind of test problem concerns the response of a thick uniform ring to sudden
internal pressure. We have developed an analytic solution to the linear problem by
standard methods, and conrmed that the nite-element solution is indeed uniform and
consistent with the analytic solution. The real purpose of ring calculations, however, is
verication that the damage is homogeneous around the ring when SCRAM is activated.
This is far from trivial for a number of reasons. First, the nite-element programs are
noisy. Second, there may be bugs and/or noise in the SCRAM algorithm or its
implementation. Third, congurations with only a few crack orientations are anisotropic
and this anisotropy leads to inhomogeneous behavior in a geometry that is nominally
symmetric. Calculations of explosions in oil shale demonstrated the utility of this approach
to verication (Dienes, 1981). Calculations with large numbers of orientations test for bugs
but also determine how many orientations are needed to get accurate damage response. We
have found that 3 orientations gives very erratic results, 9 orientations is not enough, 30 is
adequate, and 480 gives good uniformity. This conclusion, which is illustrated in Fig. 3,
requires some elaboration, however. (Note that work on related approaches typically
involves 5 or fewer orientations). The average behavior is good even when the number of
crack orientations is small. The problem is that uctuations about the average can be large
when the number of orientations is small, say 9. These calculations allow us to examine the
effect of radial cracks, which we believe is very large, and may dominate material response.
However, radial crack formation is not directly relevant to sensitivity, which depends on
the response of closed cracks. An important reason for carrying out this verication and
validation at this point is discussed in what follows.
An essential difculty with the original version of SCRAM is that there was no
provision for displaying explicitly the large cracks that are visible to the naked eye; nor did
the large cracks formed by coalescence dominate failure, as they do in real catastrophes.
The concern was to account for the growth and coalescence of microcracks in a statistical
manner. This approach was adequate at modest levels of damage, and we computed
coalescence in order to account for the limited damage that occurs when microcracks
intersect and is a feature of brittle failure. More extensive damage (rupture) frequently
occurs when macroscopic cracks form from microcracks via a percolation process.
(Cleavage is a competing mechanism.) In a sense the percolation phenomenon was implicit
in the original SCRAM, but there was no mechanism for determining the set size that is a
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1257
consequence of many intersections. An approach to computing the set size and the
percolation threshold has been under development for several years, but this process is
rather subtle and we were reluctant to incorporate it into SCRAM, which is already
complex, without extensive testing. Hence the emphasis on verication and validation
discussed above.
ARTICLE IN PRESS
Fig. 3. A comparison of the response of a thick ring to internal pressure for 4 levels of resolution of crack
orientation. With only 3 orientations the anisotropy is very large, while with 480 the anisotropy is small and the
uniformity of the response is good. For many purposes 30 orientations would give satisfactory results. The colors
denote the specic energy consumed in brittle fracture ranging from 0 (blue) to 6 10
5
erg=g (red). Lengths are in
centimeters. The internal pressure was ramped to 25 bar in 30 ms, after which it remained constant. The response
does not change after the time shown, 200 ms.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1258
3.1. Validation of the melting and ignition algorithm
In the rst analyses of crack heating as an initiation mechanism, it was assumed that
ignition could be modeled by solving the heat equation with an Arrhenius source term and
a boundary condition representing the heat generation due to interfacial friction in closed
penny-shaped cracks. As the theory was rened there was some concern that melting might
play a role in reducing the heat generated when the crack surface reaches the melting point.
On further analysis it appeared that the molten boundary layer at the crack surface is so
thin that signicant heat is produced by shearing of the viscous melt, especially when the
crack is unstable, but the latent heat of melting was accounted for only at the crack
surface. More recently, we have rened the treatment of latent heat. To investigate its
effect, we have developed a one-dimensional nite-difference ignition model containing a
cell between the molten and solid regions that accounts for discontinuous heat ux.
Motion of this special cell accounts for the latent heat of melting. The algorithm has been
validated in several ways. First, when the heat ux is held constant at the boundary the
temperature rises according to the square root of time, as predicted by a well-known
analytic solution. Second, the time to ignition predicted by the current model in a semi-
innite medium has been veried by comparison with results of Cook (1958), Linan and
Williams (1971), and Dienes (1995). Our prediction is within a few percent of the time
predicted using the theory of Linan and Williams. Third, the energy balance accounting for
heat input, viscous dissipation, reaction, and melting, is good to a few percent. Fourth, the
results are insensitive to changes in resolution when Dx
2
/Dt is xed. For the semi-innite
problem when the temperature at a boundary is xed (above the melting point),
the numerical solution is compared with the analytic (similarity) solution available in the
literature (Carslaw and Jaeger, 1959). For this problem, the numerical solution shows the
melting front moving away from the boundary with distance increasing with the square
root of time, as predicted by the analytic solution. The numerical and analytic solutions
are compared in Fig. 4, showing agreement to within a few percent at a typical time.
The precision could be increased by rening the numerical algorithm, but this does not
seem justied at this point since few of the required material properties are known to better
than a few percent accuracy, and they are not, in fact, constant. Furthermore, the
assumption of one-dimensional heat ow due to friction in a penny-shaped crack is
somewhat idealized.
In a typical initiation calculation the effect of latent heat was to increase the pre-ignition
time 34%, from 139 to 187 ms. This is the time when the Arrhenius source term has
contributed an amount of heat equal to 1% of the total heating. The actual latent heat was
taken as 2:081 10
9
erg=g, while the lowered (and negligible) value was 2:08 10
8
erg=g.
In that calculation, the rate of heating was _ q = 1:0 10
11
erg=cm
2
=s until the melting point
(T
m
= 519 K) was reached. After melting, viscous heating due to an assumed sliding
velocity of 0:35 cm=ms produced a fairly intense source term because the molten layer was
very thin (7:94 mm at ignition for the estimated latent heat, 9:03 mm for the lowered value).
A similar result may be obtained for the problem with the xed temperature at a
boundary. In that case, when the surface temperature is xed at 700 K the melt layer moves
0:42 mm with latent heat accounted for, while it moves 0:53 mm (26% more) when the latent
heat is reduced tenfold, at a time earlier than that for Fig. 4.
The treatment appears to be exible and sufciently accurate. Thus, a second
discontinuity representing the betadelta transition can be implemented when sufcient
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1259
data become available to represent the chemical-source term and the physical properties in
the two phases separately.
4. Initiation of PBX 9501
A class of high explosives (HE) known as Plastic Bonded Explosives (PBX) is widely
used in both conventional and defense applications, due to their reliable performance and
maximum safety envelopes (Thompson et al., 2002). We are concerned with modeling the
initiation of PBXs, in particular, PBX 9501, under low-amplitude mechanical impacts. As
discussed earlier, this explosive can initiate at impact velocities below 50 m/s (Idar et al.,
1998). This particular explosive is a heterogeneous material of 95% (weight percent)
energetic HMX (High Melting point eXplosive octahydro-1,3,5,7-tetranitro-1,2,3,5-
tetrazocine) grains embedded in a polymer binder, largely estane (a polyurethane), and
a small amount (0:1%) of stabilizer. The HMX is very brittle with very low specic
surface energy (about 50 erg=cm
2
). The polymer binder is viscous and much tougher. The
size of the HMX grains is random and roughly follows a bi-modal distribution with 3:1
ratio of coarse grains (average size of 200 mm) to ne ones (average size of 1 mm)
(Clements and Mas, 2004). The details of the compositions of PBX 9501, such as the
composition of the binder and the size distribution of the HMX grains, are available in the
literature (e.g., Thompson et al., 2002). Fig. 5a shows the heterogeneous microstructure
(HMX grains and polymer binder) and microcracks present in a pristine sample of PBX
ARTICLE IN PRESS
300
350
400
450
500
550
600
650
700
0 2 4 6 8 10 12 14 16 18 20
Distance (Microns)
T
e
m
p
e
r
a
t
u
r
e

(
d
e
g
r
e
e

K
e
l
v
i
n
)
2.08E9
Latent heat = 2.08E8
Fig. 4. A comparison of the effect of latent heat on thermal proles normal to a shear crack surface according to
analytic and numerical solutions with a xed temperature (700 K) on the crack surface (left boundary). The melt
front (T = 519 K) moves from left to right, its distance from the crack surface increasing with the square root of
time. The analytic solutions lie slightly above the numerical ones. The actual latent heat of PBX 9501 is
2:08 10
9
erg=g.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1260
9501 pressed to a density of 1:8 g=cm
3
(Skidmore et al., 1997), as well as (5b) the
macroscopic radial and circumferential cracks caused by a mechanical impact (Idar et al.,
1998). It also shows (5c) cracks, melt, probable reaction, and bubbles of gas products in a
fragment recovered from a quenched burn test, and (5d) a grain of HMX sheared during a
quenched test. Due to the complex microstructure, the thermo-mechanical and chemical
behaviors of this explosive are very complicated (e.g., Dienes and Kershner, 1998, 2001;
Bennett et al., 1998; Hackett and Bennett, 2000; Knauss and Sundaram, 2004).
The majority of the SCRAM model constants for PBX 9501 were obtained by tting the
model to the uniaxial compression test obtained data by Wiegand (Wiegand, 1998a, b;
Aidun, 1998). The data are taken at a strain rate of 0.01/s and at room temperature. Using
the same set of model constants, we predicted the stressstrain responses at a higher strain
rate (1000/s) and compared the predictions with the available experimental data (Dobratz
and Crawford, 1974; Funk et al., 1996). A comparison of the predicted and measured
responses is shown in Fig. 6 (the predicted response at 0.022/s is not shown because it is
essentially the same as 0.01/s). The agreement is fairly good. The stressstrain data involve
3 main features: peak stress, the corresponding strain, and the stress at 4%. By matching
these we have been able to estimate 3 microstructural material parameters: effective surface
ARTICLE IN PRESS
Fig. 5. Micrographs showing the microstructure (HMX grains and polymer binder) and cracks in a PBX 9501
(Courtesy of Skidmore and Idar of Los Alamos National Laboratory). (a) a pristine sample of PBX 9501 pressed
to a density of 1:8 g=cm
3
showing distribution of microcracks prior to testing; (b) a disk of explosive (with 5
//
diameter) impacted by a hemispherical-nosed steel projectile at low-speed (a modied Steven test by Idar et al.,
1998), illustrating brittle fracture with both radial and circumferential cracking (the projectile is slightly off
center). The cover plate has been removed to view the explosive; (c) a fragment recovered from a quenched test
showing cracks, melt, probable reaction, and bubbles of gas products and (d) a grain of HMX sheared during a
quenched test.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1261
energy ( g), number density of cracks L
0
, and the initial average size of cracks ( c
0
). It seems
remarkable that the process of tting the macroscopic data leads to a unique and plausible
set of parameters characterizing the microstructures. If this approach can be conrmed,
this would be a very useful means of characterizing the defects and failure behavior. In
Fig. 6, some adjustment of the other properties such as crack speed and elastic modulus is
made to optimize the t.
The multiple-shock experiment by Mulford et al. (1993), which involves the initiation
and quenching of the hot spots in a PBX 9501 slab, was selected for validation of both the
mechanics and initiation and burn algorithms in SCRAM. The experiment involved a two-
material yer plate, with Kel-F (Plexiglas) striking rst, followed by Vistal (Al
2
O
3
),
impacting a PBX 9501 target plate at 0:0911 cm=ms and resulting in two shocks; only the
second one is strong enough to initiate a reaction causing an observable increase in particle
velocity. The experimental setup is shown schematically in Fig. 7a. The thicknesses of
Vistal, Kel-F, and PBX 9501 are 1.1, 0.08 and 1.0 cm, respectively. Waves generated are
one-dimensional for about 4 ms. We use HYDROX (Shaw and Straub, 1981), a one-
dimensional shock physics code, with SCRAM embedded, to simulate the experiment. The
Vistal and Kel-F in the yer plate are represented using the standard HYDROX material
model, while the PBX 9501 target is modeled with SCRAM. Some model constants for the
3 materials are listed in Table 2. The properties of PBX 9501 are the same used in tting
the uniaxial stressstrain data shown in Fig. 6. In this calculation, the number of discrete
crack orientations (bins) N is chosen to be 9. The rst three have normals in the impact
direction and perpendicular to it, forming an orthogonal basis. The remaining six are
aligned with the face-diagonals of a unit cube having the orthogonal basis as edges. Thus,
under compressive shock loading, four crack bins that are at 451 to the impact direction
are under pressure and shear. Our calculation indicates that those shear cracks become
unstable and grow in size following the rst shock, becoming hot spots. The heat ux
generated by the interfacial friction on these shear cracks continues to cook the explosive,
ARTICLE IN PRESS
Fig. 6. A comparison of the stressstrain responses calculated by SCRAM at two strain rates with data obtained
from Wiegand (1998a), Funk et al. (1996), and Dobratz and Crawford (1974). The calculated response at 0.022/s
(not shown here) is essentially the same as 0.01/s.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1262
raising its temperature with time, in turn increasing the rate of chemical reaction. Then
shortly after the second shock these reacting shear cracks begin to burn as a result of
interfacial friction and heating. During burning, the gaseous product is rst highly
ARTICLE IN PRESS
Fig. 7. A comparison of the measured and calculated particle velocities for the multiple-shock impact experiment
reported by Mulford et al. (1993) in which a slab of PBX 9501 is impacted by a two-material (Kel-F and Vistal)
yer plate at 0:0911 cm=ms. The experimental setup and a time-position diagram showing intersection of shock
waves are shown in (a). The velocity plots shown in (b) are measured with 11 velocity gauges embedded in the
PBX 9501. The plots shown in (c) are calculated by HYDROX-SCRAM. The reactive heating is based on cracks
at 451 to the impact direction. The rst shock is caused the Kel-F (Plexiglas) while the second shock is due to
reection by the high-density Vistal sheet bonded to Kel-F yer. The second shock leads to frictional heating in
shear cracks sufcient to raise the local temperature and pressure, initiate reaction, and thereby increase the
material velocity.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1263
compressed inside a shear crack, developing high internal pressure which gives rise to the
observed increase in particle velocities. For an individual crack, opening and fast growth
(at roughly the sound speed) reduce the gas pressure so quickly that the crack becomes
stable, quenching the burn. We believe that this kind of behavior is responsible for the
observed initiation and subsequent quenching of burn and reactions in PBX 9501.
A comparison of the velocity histories of particles at eleven locations inside the target
calculated by HYDROX-SCRAM and measured by embedded gauges (Mulford et al.,
1993) is shown in Figs. 7b (measured) and 7c (calculated). The rst shock (jump in particle
velocities) is caused by the Kel-F while the second shock is due to reection by the high-
density Vistal sheet. The second shock leads to frictional heating in shear cracks sufcient
to raise the local temperature and pressure, initiate reaction, and thereby increase the
particle velocities. The decrease in particle velocities at late times is the result of quenching.
The agreement between the calculation and data is reasonably good, but improvement can
be hoped for, especially at the late times when our modeling of the gaseous reaction
products can be improved.
In this calculation isolated unstable cracks are ultimately quenched; however, in other
applications, interactions between cracks may exceed a percolation threshold generating a
ARTICLE IN PRESS
Table 2
Material constants (cgs units, except as noted; blank entries indicate data not needed in the calculation)
Vistal Kel-F PBX 9501
r, mass density 4.103 2.12 1.86
K, bulk modulus
4:64 10
12
80:6 10
9
93:85 10
9
S, slope of u
s
u
p
line in EOS
a
2.0 1.87 2.26
G, Gruneisens ratio in EOS
a
1.50
G, shear modulus
29:0 10
9
n, Poissons ratio 0.36
c
0
, initial crack size
32 10
4
L
0
, crack number density per 2p
45 10
3
g, specic surface energy 50.0
m
s
, static coefcient of friction 0.8
m
d
, dynamic coefcient of friction 0.2
n, power-law parameter for slow crack growth 6.0
a, coalescence parameter 3.4
Y, yield stress in pure shear
4:0 10
9
b, kinematic hardening modulus
0:2 10
9
C
v
, heat capacity
9:96 10
6
k, thermal conductivity (cal cm
1
s
1
K
1
) 9:7 10
4
T
m
, melting point (K) 519
T
i
, initial temperature (K) 300
L, latent heat of melting
2:081 10
9
E
A
, activation energy in Arrhenius kinetics (cal/mole)
5:27 10
4
Q
r
, heat of reaction (cal/g) 500
Z, pre-exponential factor in Arrhenius kinetics
5:9 10
19
g, ratio of specic heats for the reactive product 1.3
a
S and G are the material constants used in a MieGruneisen equation of state (EOS), which relates pressure to
the density and internal energy (Dienes, 1985). A nonlinear EOS is needed to capture formation of shock waves at
high pressure.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1264
dynamic instability, which can lead to a violent explosion. This behavior is too complex for
detailed modeling at this point because of the multitude of physical processes involved and
the difculty of dealing with high-speed crack growth numerically. (Still, it may be
approximated in numerical codes by cell interactions.) Dienes and Kershner (2001) model
the response of a burning crack to an oscillating pressure eld that idealizes the effect of
neighboring cracks. It would be very useful to test the mechanisms governing the late
stages of reaction; one approach is described in Section 5.
4.1. Role of crack orientation
In a recent APS symposium it was shown that the orientation of shear cracks is crucial
(Dienes et al., 2004); thus, we regard with some skepticism brittle-failure models in which
crack orientation is not carefully accounted for. In fact, new theoretical work has shown
that friction plays a major role in fracture, and determines the stress states that lead to
instability in penny-shaped cracks (Zuo and Dienes, 2005). For materials with high
coefcients of friction, the range of unstable crack orientations in compression turns out to
be rather narrow. The fundamental theory implicit in our computational model has not
needed modicationwe simply had not examined the consequences of the computational
model in detail until recently.
In the analysis of the experiment of Mulford et al. discussed above, for which the results
are given in Fig. 7, cracks in four bins (out of nine total) that had normals oriented at
451 to the impact direction were found to form frictional hot spots; the remaining ve
crack bins had normals either parallel or perpendicular to the impact direction and hence
did not shear. In the light of our recent studies of the effect of friction which suggest that
cracks more nearly aligned with the direction of impact are more critical, we repeated the
HYDROX-SCRAM calculation of the experiment using several different crack orienta-
tions for the four crack bins which formed hot spots; the orientations of the remaining ve
bins were unchanged. The effect of crack orientation is shown in Fig. 8, where the angle is
that between the crack normals and the impact direction and was kept the same for all four
crack bins during the calculations. We conclude that the most critical crack has its normal
near 551 to the impact direction (an angle of 351 between the crack plane and the impact
direction), consistent with our theoretical prediction.
4.2. Effects of the viscosity
As with most modern materials models, a variety of material properties are required to
carry out realistic SCRAM calculations. Some can be obtained by molecular dynamics
(MD) simulation, but properties that depend on the history of the material sample and its
resident defects, such as porosity and ultimate strength, can only be obtained from direct
measurements. When properties depend only on molecular structure and can be obtained
by MD, we have an advantage in that the effects of temperature and pressure can be
accounted for. In this paper only one example will be considered, the viscosity of HMX,
for which calculations have been carried out by Bedrov et al. (2000) (BSS). A strong
sensitivity to temperature was found. The values of viscosity rst used in our calculations
result in relatively modest heating in the shear layer, but the values in BSS allow our
calculated velocities to agree reasonably well with the very precise multiple shock
experiment of Mulford et al., as shown in Fig. 7 and discussed above.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1265
The heating due to shear in HMX has been considered by Winter and Field (1975), Frey
(1981), and Dienes and Kershner (1998). Values of viscosity based on the reports of these
authors vary from below 0.001 to 0.0139 Pa s (MKS units are used throughout this section
to maintain consistency with BSS). BSS have actually calculated the viscosity by means of
MD, and obtained the result:
m(T) = m
0
exp(E
vis
=(RT)) (64)
with parameters E
vis
, m
0
determined by tting the formula with the viscosities at 800 and
750 K calculated by MD: m
0
= 3:45 10
7
Pa s; E
vis
=R = 7744 K (a misprint in BSS says
m(T) = m
0
exp(E
vis
=(RT))). The viscosity used in the original calculations of Dienes and
Kershner (0.00726) corresponds to Eq. (64) at 778 K. However, at temperatures near the
melting point (519 K) the formula gives much higher viscosity.
Plots of the computed particle velocities for the experiment by Mulford et al. (1993) are
compared in Fig. 9 for various viscosities. With the constant value of 0.0139, the upper
ARTICLE IN PRESS
Fig. 8. The effect of crack orientation on the computed response of PBX 9501 in the experiment of Mulford et al.
(1993) described in the text and in Fig. 7. Time is in microseconds, and velocity is in cm=ms. (a) 451 (same as Fig.
7c); (b) 551; (c) 701; (d) 801. The angle is that between the crack normal and the direction of impact. Only crack
orientations in a small range lead to signicant reactions. (The velocity remains high at late times because we have
not allowed for a transition from solid to burn products in the solid EOS.)
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1266
ARTICLE IN PRESS
Fig. 9. The effect of melt viscosity: (a) m = 0:0139Pa s; (b) m = 1:042 and (c) m(T) = 3:45 10
7
exp(7744=T).
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1267
limit of the reported values, there is no ignition (Dienes, 1996), as shown in Fig. 9a. Taking
the viscosity at 600 K from Eq. (64), 0.139, there is still no signicant reaction (not shown).
With the constant viscosity at 550 K (the lower limit of the calculations of BSS), 0.45, there
is some reaction, but the peak velocity is well below the measured value. Raising the
viscosity to 1.042, the value at the melting point obtained from the BSS formula (which
extrapolates the BSS calculations) increases the peak velocity, but it is still somewhat
below the measured value, as shown in Fig. 9b. The calculation using the temperature-
dependent viscosity given by Eq. (64) gives the best result (Fig. 9c). It is emphasized that in
addition to the temperature-dependency, the values of viscosity calculated from Eq. (64)
are signicantly higher than the values used in previous studies (Winter and Field, 1975;
Frey, 1981; Dienes and Kershner, 1998), especially at temperatures near the melting point.
This comparison illustrates the importance of good estimates of the temperature
dependence of material properties, particularly if we are to compute the effects of
localization and its effect on explosive hot spots. A particular challenge is the coupling of
chemical, thermal and mechanical properties in hot spots as they transition toward
explosive reactions. This requires experimental guidance because of the complexity of the
combined chemical, physical, and mechanical processes, as discussed in the section that
follows.
5. Discussions and recommendations
Nonlinear behavior of some materials can be idealized with plasticity theory or some
generalization thereof, but explosives and propellants exhibit much more complex
behavior which should be accounted for in assessing the risk of accidental explosions.
In particular, most ductile materials such as metals do not show signicant size effects, but
explosives are typically sensitive to the scale of the mass being investigated, as are most
solids in a brittle condition. In particular, larger samples generally have lower strength
because they contain larger defects, but also a higher probability that smaller defects will
combine (Bazant and Planas, 1998; Persson et al., 1994). A related complication is that
strength may uctuate between samples, again due to the presence of defects. Failure of
brittle materials may be catastrophic, with the speed of crack propagation near or even
exceeding the shear wave speed in special cases (Rosakis et al., 1999; Rosakis, 2002). While
brittle materials may behave stably in compression and appear to have a certain ductility,
they may exhibit sudden failure. The materials used in explosives and propellants often
have brittle constituents, though they appear to be plastic when taken in a large mass due
to a soft binder material. The test results shown in Fig. 6 shows that signicant failure may
occur at only 2% strain in one common explosive, and incipient failure occurs at an even
lower strain. SCRAM theory provides a framework for dealing with some of these
complexities, but it needs to be combined with appropriate experiments to make reliable
predictions of risk possible. Experiments are necessary because the defect structure
depends on the actual history of material which is rarely known and, consequently,
requires experiments to characterize it.
To illustrate, consider the XDT data cited in the Introduction, where 12 in 50 of the
impact tests of propellant cylinders resulted in a violent reaction, while the others exhibited
a mild deagration. Impact analysis with SCRAM has shown that the region suffering
signicant frictional heating lies in a rectangular torus with a 1 1 mm
2
section and a
mean radius of 2 mm, leading to a critical volume of 0:012 cm
3
. The number density of
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1268
grains was measured at 0:85 10
6
=cm
3
and the average grain size was 100 mm. We estimate
the typical number density of large cracks (about 1% of the number density of grains) as
1:0 10
4
=cm
3
and the average radius of defects, when idealized as penny-shaped cracks, as
10 mm. With an exponential distribution of crack sizes we can conclude that the number
density of cracks exceeding 50 mm in radius is 67=cm
3
. We assume that 1/4 of these have an
orientation that would make them unstable in compression. Then the expected number of
dangerous cracks per shot would be 0:012 67=4 = 0:2, i.e., roughly 1 in 5 explodes, as
observed. This estimate cannot be taken too seriously because the details of the reactive
process are not fully established; the point of the calculation is to illustrate that combining
defect behavior with statistics can lead to useful estimates of risk, and that sample size is
crucial.
It has long been known that the sensitivity of explosives is affected by their porosity,
even when below 1%, but the actual hot spot mechanism has never been conrmed
experimentally. At high pressures some investigators have attributed the sensitivity to the
collapse of spherical hot spots, but this mechanism does not explain initiation at low
impact speeds such as those studied by Jensen et al. (300 m=s) or Idar et al. (50 m=s).
More specically, it has not been established whether hot spots are initiated within the
binder, at its interface with an explosive grain, or within the explosive grain itself. As
discussed in Section 4, HMX is very brittle (with specic surface energy about 50 erg=cm
2
),
hence we take intragranular fracture to be the dominant mechanism. This view is
supported by recent experimental evidence concluding that explosive sensitivity is a result
of intragranular, not extragranular, voids (Borne and Beaucamp, 2002). This does not of
itself conrm that shear cracks dominate initiation at low speeds, but it is supportive.
Though cracks may be initially open, they can be quite thin, so that a modest pressure
could close them and the subsequent behavior would be that of a shear crack. Recent
studies show that unstable shear cracks can propagate at above sonic speeds (Rosakis,
2002; Abraham, 2001; Gao et al., 2001). Melting may well occur due to interfacial friction
in such cracks since the interfacial sliding velocity is related to the crack growth rate; the
formation of pseudotachyltes, which is being actively explored in the geophysics literature
(Scholz, 2002; Di Toro et al., 2005), is evidence that melting occurs in faults. Explosives
have a much lower melting point than rocks so they can melt easily under intense frictional
heating. In many explosives melting is accompanied by reactions that form gaseous
products that could open the shear cracks, possibly inducing further instability. Crack
interactions result from both stress waves (Dienes and Kershner, 2001) and intersections.
(Permeability due to crack intersections can be important in explosives because it would
allow hot reaction products to be more readily transported within the solid than by
diffusion.) Furthermore, the interaction of hot spots is probably crucial to initiation, but is
difcult to assess. However, experimental evidence that hot-spots will die out without
interactions was given recently by Proud et al. (2002, 2004) and Field et al. (2004).
The need for better insights is clear from the results of drop tests and such, where a
number of repetitions are needed and an average is noted, though the scatter may be very
large. As pointed out by Chakravarty et al. (2002), ordering of explosives by sensitivity
cannot be done. Different tests provide different ordering. Furthermore, standard tests do
not show by what process porosity, grain size, and method of fabrication inuence
sensitivity. As restrictions on testing increase for reasons of safety and economics, it
becomes more attractive to make use of theoretical simulations, but the details of damage,
initiation, and variability need to be better understood before we can rely on computer
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1269
simulations. Though we have made idealized calculations of shear crack heating, dynamic
three-dimensional shear crack behavior may be much more complex, and the critical crack
size and orientation for a specic scenario may be different from our predictions.
In the SCRAM formalisms, we have attempted to quantify these issues in the simplest
way possible, retaining the mechanisms that dominate brittle behavior. In this paper, we
emphasize the behavior of reactive materials such as explosives and propellants, but the
formalism has been used to study other kinds of brittle behavior as well, which has provided
additional validation (Meyer et al., 1999; Zuo et al., 2003). Unfortunately, there have been
no experimental programs designed to provide detailed information on sensitivity and crack
statistics in the same material, allowing us to conrm that the shear-crack mechanism of
initiation dominates sensitivity; we can only claim that SCRAM results are consistent with
observations. In particular, only a few mechanical properties of explosive materials are
normally reported, and such data as the mean crack size, number density of cracks, friction
coefcient, fracture toughness, and shear modulus must be inferred.
How can assessments of risk be conrmed? Even if extensive defect and reaction data
were available, there could remain some concerns that the mechanisms for initiation might
be overly simplied. A natural validation procedure would be to examine samples of
various sizes to see that the trend is as predicted. This can be expensive, and such programs
have had limited success in the eld of rock mechanics. An alternative would be to create a
mixture of explosive and sugar (or other inert that simulates explosive grains, what we call
a PBMIX) and examine the damage and sensitivity to impact when the fraction of
explosive is increased, as illustrated in Fig. 10. The extent of fracture, reaction, and phase
changes in the explosive grains could be safely examined following an impact near critical
conditions, while test samples on a realistic explosive device near threshold conditions
cannot be safely recovered and examined. (The tests of Howe et al. (1985) with polishing of
damaged TNT could probably not be repeated today.)
Specic results from tests of a PBMIX that would not be available from standard
experiments are listed below:
+ Severely damaged HMX grains can be recovered and examined as part of a nearly intact
sample.
+ The location, orientation, and shape of fractures and hot spots can be determined.
+ Large pieces can be conveniently sectioned, polished, etched, and examined for phase
changes and reactions at grain and crack surfaces.
ARTICLE IN PRESS
Fig. 10. Conceptual diagram of the expected response from testing a PBMIX. The red region represents largely
reacted material, the dark blue dots denote partially reacted material (the zz zone), pink denotes unreacted crack
surface, and green denotes binder.
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1270
+ The importance of hot-spot interactions can be assessed. (Are adjacent hot spots hotter
and more extensively reacted than isolated ones?)
+ Crack and hot-spot statistics can be obtained and correlated. The number and size of
defects that are stable, exhibit phase changes, and show evidence of reaction can be
compared.
+ Physically motivated models for the initiation of burning and hot-spot interactions can
be validated.
6. Summary and conclusions
We have presented a statistical theory (SCRAM) for modeling impact initiation of
reactive materials (explosives and propellants) which contain brittle constituents. The
theory has been implemented and was used to model a multiple-shock experiment
on a typical explosive (PBX 9501). The calculated particle velocities compare reasonably
well with those measured using gauges embedded inside the explosive. The effects of
crack orientation and temperature dependence of viscosity of the melt on the response
have also been examined. It is shown that crack orientation has a signicant effect on the
response, especially under compressive loading where interfacial friction plays an
important role.
Our hypothesis for the initiation mechanism is that intense frictional heating in a shear
crack during unstable growth forms a hot spot, and that the reactive material adjacent to a
hot spot can undergo localized melting, ignition, and fast burn, which can lead to violent
explosions following relatively mild impacts. The evidence for this hypothesis has been
discussed and it appears to be a viable explanation for a variety of events. Idealization of
hot spots as penny-shaped cracks makes it possible to model individual crack behavior, as
well as ensembles of cracks, theoretically. The behavior of open cracks and closed cracks
with friction are quite different, but the model lends itself to detailed analysis of both. In
addition it is feasible to examine the interactions of defects in the form of circular discs and
to explain how ensembles of defects may form violent reactions, such as XDT. We have
shown (Dienes, 2005) that an ensemble of circular cracks has a percolation threshold such
that a network of connected cracks forms above that threshold.
Shear cracks are hard to observe in any detail because there is so little of them, and the
heating they produce is hard to observe because it is so transient. Means for observing the
number and interactions of shear cracks and hot spots in more detail should be developed
in order to conrm quantitative estimates of risk and size effects. Tests of a mixture of
reactive and inert materials, a PBMIX, would make it possible to quantify the evolution of
individual hot spots in an ambiance of reduced violence.
Acknowledgments
This work was supported by the Joint US Department of Energy (DOE) and US
Department of Defense (DoD) Munitions Technology Development Program, the DOE
Advanced Simulation and Computing (ASC) Program. We are indebted to D.A. Wiegand
and J.B. Aidun for providing the experimental stressstrain data for PBX 9501, which were
used to determine the model constants. We thank J. Middleditch for his contributions to
the model implementation and calculations, and C.A. Bronkhorst, R.M. Hackett, C. Liu
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1271
for technical discussions. We are particularly grateful to W.G. Proud for his thorough
reading of an earlier version of the manuscript and providing many constructive
comments.
References
ABAQUS, 1998. ABAQUS Theory Manual, Version 5.8. Hibbitt, Karlsson & Sorenson, Inc., Pawtucket, RI.
Abraham, F.F., 2001. The atomic dynamics of fracture. J. Mech. Phys. Solids 49, 20952111.
Addessio, F.L., Johnson, J.N., 1990. A constitutive model for the dynamic response of brittle materials. J. Appl.
Phys. 67, 32753286.
Aidun, J.B., 1998. Private communication.
Antoun, T., Seaman, L., Curran, D.R., Kanel, G.I., Razorenov, S.V., Utkin, A.V., 2003. Spall Fracture. Springer,
New York.
Asay, B.W., Laabs, G.W., Henson, B.F., Funk, D.J., 1997. Speckle photography during dynamic impact of an
energetic material using laser-induced uorescence. J. Appl. Phys. 10931099.
Balzer, J.E., Field, J.E., Gifford, M.J., Proud, W.G., Walley, S.M., 2002. High-speed photographic study of the
drop-weight impact response of ultrane and conventional PETN and RDX. Combustion Flame 130,
298306.
Bazant, Z.P., Planas, J., 1998. Fracture and Size Effect in Concrete and Other Quasibrittle Materials. CRC Press,
Boca Raton.
Bedrov, D., Smith, G.D., Sewell, T.D., 2000. Temperature-dependent shear viscosity coefcient of octahydro-
1,3,5,7-tetranitro-1-3-5-7-tetrazocine (HMX): a molecular dynamics simulation study. J. Chem. Phys. 112,
72037208.
Belytschko, T., Liu, W.K., Moran, B., 2000. Nonlinear Finite Elements for Continua and Structures. Wiley, New
York.
Bennett, J.G., Haberman, K.S., Johnson, J.N., Asay, B.W., Henson, B.F., 1998. A constitutive model for the non-
shock ignition and mechanical response of high explosives. J. Mech. Phys. Solids 46, 23032322.
Bonnett, D.L., Butler, P.B., 1996. Hot-spot ignition of condensed energetic materials. J. Propulsion Power 12,
680690.
Borne, L., Beaucamp, A., 2002. Effects of explosive crystal internal defects on projectile impact initiation. In:
Short, J.M. (Ed.), Proceedings of the 12th International Detonation Symposium. The Ofce of Naval
Research, Arlington, VA, pp. 3543.
Bowden, F.P., Yoffe, Y.D., 1952. Initiation and Growth of Explosion in Liquids and Solids. Cambridge
University Press, UK.
Carslaw, H.S., Jaeger, J.C., 1959. Conduction of Heat in Solids, second ed. Clarendon Press, Oxford, UK.
Charles, R.J., 1958. Dynamic fatigue of glass. J. Appl. Phys. 29, 16521657.
Chakravarty, A., Gifford, M.J., Greenway, M.W., Proud, W.G., Field, J.E., 2002. Factors affecting shock
sensitivity of energetic materials. In: Furnish, M.D., Thadhani, N.N., Horie, Y. (Eds.), Shock Compression of
Condensed Matter2001, AIP Conference Proceedings, vol. 620. AIP Press, New York, pp. 10071010.
Chaudri, M.M., 1972. Shock initiation of fast decomposition in crystalline solids. Combustion Flame 19, 419425.
Clements, B.E., Mas, E.M., 2004. A theory for plastic-bonded materials with a bimodal size distribution of ller
particles. Model. Simul. Mater. Sci. Eng. 12, 407421.
Cook, G.B., 1958. The initiation of explosion in solid secondary explosives. Proc. R. Soc. London A 246, 154160.
Crank, J., 1984. Free and Moving Boundary Value Problems. Clarendon Press, Oxford, UK.
Curran, D.R., Seaman, L., 1996. Simplied models of fracture and fragmentation. In: Davison, L., Grady, D.E.,
Shahinpoor, M. (Eds.), High Pressure Shock Compression of Solids II. Springer, New York, pp. 340365.
Curran, D.R., Seaman, L., Shockey, D.A., 1987. Dynamic failure of solids. Phys. Report 147, 253288.
Curran, D.R., Seaman, L., Copper, T., Shockey, D.A., 1993. Micromechanical model for comminution and
granular ow of brittle material under high strain rate application to penetration of ceramic targets. Int. J.
Impact Eng. 13, 5383.
Dienes, J.K., 1978. A statistical theory of fragmentation. In: Kim, Y.S. (Ed.), Proceedings of 19th US Rock
Mechanics Symposium. University of Nevada, pp. 5155.
Dienes, J.K., 1979. On the analysis of rotation and stress rate in deforming bodies. Acta Mech. 32,
217232.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1272
Dienes, J.K., 1981. On the effect of anisotropy in explosive fragmentation. In: Einstein, H.H. (Ed.), Rock
Mechanics from Research to ApplicationProceedings of 22nd US Rock Mechanics Symposium. MIT,
Cambridge, MA, pp. 177183.
Dienes, J.K., 1982. A frictional hot-spot theory for propellant sensitivity. In: Proceedings of the Second
JANNAF Propulsion Systems Hazards Subcommittee Meeting. Naval Weapons Center, China Lake, CA,
pp. 269278.
Dienes, J.K., 1984. Frictional hot-spot and propellant sensitivity. In: Proceedings of the Materials Research
Society Symposium, vol. 24, pp. 373381.
Dienes, J.K., 1985. A statistical theory of fragmentation processes. Mech. Mater. 4, 325335.
Dienes, J.K., 1986. On reactive shear bands. Phys. Lett. A 118, 433438.
Dienes, J.K., 1989. Theory of Deformation, part II, Physical Theory. Los Alamos National Laboratory Report
LA-11063-MS vol. II.
Dienes, J.K., 1995. Verifying and understanding the friction-ignition process. Personal communication to P.M.
Howe.
Dienes, J.K., 1996. A unied theory of ow, hot spots, and fragmentation with an application to explosive
sensitivity. In: Davison, L., Grady, D.E., Shahinpoor, M. (Eds.), High Pressure Shock Compression of Solids
II. Springer, New York, pp. 366398.
Dienes, J.K., 2001. Crack dynamics via Lagranges equations and generalized coordinates. Acta Mech. 148,
7992.
Dienes, J.K., 2003. Finite deformation of materials with an ensemble of defects. Los Alamos National Laboratory
Report LA-13994-MS.
Dienes, J.K., 2005. On the mean cluster size in a network of cracks. Special Issue of Structural Control and Health
Monitoring, honoring the memory of Prof. T.K. Caughey, to be published, doi:10.1002/stc.126.
Dienes, J.K., Kershner, J.D., 1998. Multiple-shock initiation via statistical crack mechanics. In: Short, J.M.,
Kennedy, J.E. (Eds.), Proceedings of 11th Detonation Symposium, pp. 717724.
Dienes, J.K., Kershner, J.D., 2001. Crack dynamics and explosive burn via generalized coordinates. J. Comput.
Aided Mater. Design 7, 217237.
Dienes, J.K., Middleditch, J., Kershner, J.D., Zuo, Q.H., Starobin, A., 2002. Progress in statistical crack
mechanics: an approach to initiation. In: Short, J.M. (Ed.), Proceedings of the 12th International Detonation
Symposium. The Ofce of Naval Research, Arlington, VA, pp. 793799.
Dienes, J.K., Middleditch, J., Zuo, Q.H., Kershner, J.D., 2004. On the role of crack orientation in brittle failure.
In: Furnish, M.D., Gupta, Y.M., Forbes, J.W. (Eds.), Shock Compression of Condensed Matter-2003, AIP
Conference Proceedings, vol. 706. Springer, New York, pp. 447450.
Di Toro, G., Nielsen, S., Pennacchioni, G., 2005. Earthquake rupture dynamics frozen in exhumed ancient faults.
Nature 436, 10091012.
Dobratz, B.M., Crawford, P.C., 1974. Properties of chemical explosives and explosive stimulants. Livermore
National Laboratory Report, UCRL-51319, Rev.1.
Espinosa, H.D., 1995. On the dynamic shear resistance of ceramic composites and its dependence on applied
multiaxial deformation. Int. J. Solids Struct. 32, 31053128.
Evans, A.G., 1974. Slow crack growth in brittle-materials under dynamic loading conditions. Int. J. Fract. 10,
251259.
Field, J.E., Bourne, N.K., Palmer, S.J.P., Walley, S.M., 1992. Hot-spot ignition mechanisms for explosives and
propellants. Philos. Trans. R. Soc. A 339, 269283.
Field, J.E., Walley, S.M., Proud, W.G., Balzer, J.E., Gifford, M.J., Grantham, S.G., Greeaway, M.W., Sivour,
C.R., 2004. The shock initiation and high strain rate mechanical characterization of ultrane energetic
powders and compositions. In: Armstrong, R., Thadhani, N., Wilson, W., Gilman, J., Simpson, R. (Ed.),
Proceedings of the Material Research Society Symposium, vol. 800, pp. 179190.
Frank-Kamenetzky, A.A., 1942. On the mathematical theory of thermal explosions. Acta Physicochim. URSS 16,
357361.
Freund, L.B., 1990. Dynamic Fracture Mechanics. Cambridge University Press, New York.
Frey, R.B., 1981. The initiation of explosive charges by rapid shear. In: Short, J.M. (Ed.), Proceedings of the
Seventh Symposium on Detonation, pp. 3642.
Funk, D.J., Laabs, G.W., Peterson, P.D., Asay, B.W., 1996. Measurement of the stress/strain response of
energetic materials as a function of strain rate and temperature: PBX 9501 and mock 9501. In: Schmidt, S.C.,
Tao, W.C. (Eds.), Shock Compression of Condensed Matter, AIP Conference Proceedings, vol. 370. AIP
Press, New York, pp. 145148.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1273
Gailly, B.A., Espinosa, H.D., 2002. Modelling of failure mode transition in ballistic penetration with a continuum
model describing microcracking and ow of pulverized media. Int. J. Numer. Methods Eng. 54, 365398.
Gao, H., Huang, Y., Abraham, F.F., 2001. Continuum and atomistic studies of intersonic crack propagation.
J. Mech. Phys. Solids 49, 21132132.
Grady, D.E., 1988. Spall strength of condensed matter. J. Mech. Phys. Solids 36, 353384.
Green, L.G., James, E., Lee, E.L., Chambers, E.S., Tarver, C.M., Westmoreland, C., Weston, A.M., Brown, B.,
1981. Delayed detonation in propellants from low velocity impact. In: Short, J.M. (Ed.), Proceedings of the
Seventh Symposium on Detonation, pp. 254264.
Gurtin, M.E., 1981. An Introduction to Continuum Mechanics. Academic Press, New York.
Hackett, R.M., Bennett, J.G., 2000. An implicit nite element material model for energetic particulate composite
materials. Int. J. Numer. Methods Eng. 49, 11911209.
Hallquist, J.O., Whirley, R.G., 1989. DYNA3D Users Manual. UCID-19592, Lawrence Livermore National
Laboratory.
Heaton, T.H., 1990. Evidence for and implications of self-healing pulses of slip in earthquake ruptures. Phys.
Earth Planet. Interiors 64, 120.
Howe, P.M., Gibbons, G.G., Webber, P.E., 1985. An experimental investigation of the role of shearing initiation
of detonation. In: Short, J.M., Deal, W.E. (Eds.), Proceedings of the Eighth Symposium on Detonation,
pp. 294306.
Idar, D.J., Lucht, R.A., Straight, J.W., Scammon, R.J., Browning, R.V., Middleditch, J., Dienes, J.K., Skidmore,
C.B., Buntain, G.A., 1998. Low amplitude insult project: PBX 9501 high explosive violent reaction
experiments. In: Short, J.M., Kennedy, J.E. (Eds.), Proceedings of the 11th Detonation Symposium,
pp. 101110.
Jensen, R.C., Blommer, E.J., Brown, B., 1981. An instrumented shotgun facility to study impact initiated
explosive reactions. In: Short, J.M. (Ed.), Proceedings of the Seventh Symposium on Detonation, pp. 299307.
Johnson, G.R., Strk, R.A., Beissel, S.R., 2001. User instructions for the 2001 version of the EPIC code. Alliant
Techsystems Incorporated, Hopkins, Minnesota.
Kanninen, M.F., Popelar, C.H., 1985. Advanced Fracture Mechanics. Oxford University Press, New York.
Keer, L.M., 1966. A note on shear and combined loading for a penny-shaped crack. J. Mech. Phys. Solids 14, 16.
Knauss, W.G., Sundaram, S., 2004. Pressure-sensitive dissipation in elastomers and its implication for the
detonation of plastic explosives. J. Appl. Phys. 96, 72547266.
Korvin, G., 1989. Fractured but not fractal: fragmentation of the gulf of Suez basement. Pure Appl. Geophys.
131, 289305.
Krajcinovic, D., 1998. Selection of damage parameterart or science? Mech. Mater. 28, 165179.
Linan, A., Williams, F.A., 1971. Theory of ignition of a reactive solid by constant energy ux. Combustion Sci.
Technol. 3, 9198.
Mader, C.L., 1979. Numerical Modeling of Detonations. University of California Press, Berkeley, CA.
Meyer, Jr., H.W., Abeln, T., Bingert, S., Bruchey, W.J., Brannon, R.M., Chhabildas, L.C., Dienes, J.K.,
Middleditch, J., 1999. Crack behavior of ballistically impacted ceramic. In: Furnish, M.D., Chhabildas, L.C.,
Hixson, R.S. (Eds.), Shock Compression of Condensed Matter, AIP Conference Proceedings, vol. 505.
Springer, New York, pp. 11091112.
Mulford, R.N., Shefeld, S.A., Alcon, R.A., 1993. Initiation of preshocked high explosives PBX-9404, PBX-9502,
PBX-9501, monitored with in-material magnetic gauging. In: Proceedings of the 10th International
Detonation Symposium, pp. 459467.
Oda, M., 1983. A method for evaluating the effect of crack geometry on the mechanical behavior of cracked rock
masses. Mech. Mater. 2, 163171.
Oda, M., Suzuki, K., Maeshibu, K., 1984. Elastic compliance for rock-like materials with random cracks. Soils
Found. 24, 2740.
Persson, P.A., Lee, J., Holmberg, R., 1994. Rock Blasting and Explosives Engineering. CRC Press, Boca Raton.
Prager, W., 1955. The theory of plasticity: a survey of recent achievements. Proc. Inst. Mech. Eng. 169, 4157.
Proud, W.G., Crossland, E.J.W., Field, J.E., 2002. High-speed photography and spectroscopy in determining the
nature, number and evolution of hot spots in energetic materials. In: Proceedings of SPIEThe International
Society for Optical Engineering: 25th International Congress on High-Speed Photography and Photonics. vol.
4948, pp. 510518.
Proud, W.G., Kirby, I.J., Field, J.E., 2004. The nature, number, and evolution of hot-spots in ammonium nitrate.
In: Furnish, M.D., Gupta, Y.M., Forbes, J.W. (Eds.), Shock Compression of Condensed Matter2003, AIP
Conference Proceedings, vol. 706. Springer, New York, pp. 10171020.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1274
Rice, J.R., 1984. Comments on On the stability of shear cracks and the calculation of compressive strength by
J.K. Dienes. J. Geophys. Res. 89, 25052507.
Rice, J.R., 2000. New perspectives on cracks and fault dynamics. In: Aref, H., Phillips, J.W. (Eds.), Mechanics for
a New Millennium: Proceedings of the ICTAM 2000. Kluwer Academic Publishers, Dordrecht, pp. 123.
Rosakis, A.J., 2002. Intersonic shear cracks and fault ruptures. Adv. Phys. 51, 11891257.
Rosakis, A.J., Samudrala, Coker, O., 1999. Cracks faster than the shear wave speed. Science 284, 13371340.
Sack, R.A., 1946. Extension of Grifths theory of rupture to three dimensions. Proc. Phys. Soc. 58, 729736.
Sammis, C.G., Biegel, R.L., 1989. Fractals, fault-gouge, and friction. Pure Appl. Geophys. 131, 255271.
Schmitt, D.R., Ahrens, T.J., Svendsen, B., 1989. Shocked-induced melting and shear banding in single crystal
NaCl. J. Geophys. Res. 94, 58515871.
Scholz, C.H., 2002. The Mechanics of Earthquakes and Faulting, second ed. Cambridge University Press, New
York.
Seaman, L., Curran, D.R., Shockey, D.A., 1976. Computational models for ductile and brittle fracture. J. Appl.
Phys. 47, 48114826.
Segedin, C.M., 1950. Note on a penny-shaped crack under shear. Proc. Cambridge Philos. Soc. 47, 396400.
Shaw, M.S., Straub, G.K., 1981. HYDROX: A one-dimensional Lagrangian hydrodynamics code. Los Alamos
National Laboratory Report LA-8642-M Manual UC-32.
Simo, J.C., Hughes, T.J.R., 1998. Computational Inelasticity. Springer, New York.
Skidmore, C.B., Phillips, D., Crane, N.B., 1997. Microscopical examination of plastic-bonded explosives.
Microscope 45, 127136.
Stroh, A.N., 1957. A theory of fracture of metals. Adv. Phys. 6, 418465.
Taylor, L.M., Flanagan, D.P., 1987. PRONTO 3D: A Three-dimensional Transit Solid Dynamics Program,
SAND 87-1912, Sandia National Laboratories.
Thompson, D.G., Idar, D.J., Gray, G.T., Blumenthal, W.R., Cady, C.M., Roemer, E.L., Wright, W.J., Peterson,
P.D., 2002. Quasi-static and dynamic mechanical properties of new and virtually-aged PBX 9501 composites
as a function of temperature and strain rate. In: Short, J.M. (Ed.), Proceedings of the 12th International
Detonation Symposium. The Ofce of Naval Research, Arlington, VA, pp. 363368.
Ward, M.J., Son, S.F., Brewster, M.Q., 1998. Steady deagration of HMX with simple kinetics: a gas phase chain
reaction model. Combustion Flame 114, 556568.
Wiegand, D.A., 1998a. Mechanical failure properties of composite plastic bonded explosives. In: Schmidt, S.C.,
Dandekar, D.P., Forbes, J.W. (Eds.), Conference of the American-Physical-Society Topical Group on Shock
Compression of Condensed Matter, AIP Conference Proceedings, vol. 429. AIP Press, New York,
pp. 599602.
Wiegand, D.A., 1998b. Private communication.
Winter, R.E., Field, J.E., 1975. The role of localized plastic ow in the impact initiation of explosives. Proc. R.
Soc. London A 343, 399413.
Zuo, Q.H., Dienes, J.K., 2005. On the stability of penny-shaped cracks with friction: the ve types of brittle
behavior. Int. J. Solids Struct. 42, 13091326.
Zuo, Q.H., Dienes, J.K., Middleditch, J., Meyer, H.W., 2003. Modeling the damage in ceramic armor via
statistical crack mechanics. In: Proceedings of the 2003 SEM Annual Conference and Exposition on
Experimental and Applied Mechanics, Society for Experimental Mechanics, Inc., CT, 9p.
ARTICLE IN PRESS
J.K. Dienes et al. / J. Mech. Phys. Solids 54 (2006) 12371275 1275

You might also like