Sie sind auf Seite 1von 9

Timothy J.

Holmquist
Gordon R. Johnson
Southwest Research Institute,
5353 Wayzata Blvd.,
Minneapolis, MN 55416

A Computational Constitutive
Model for Glass Subjected to
Large Strains, High Strain Rates
and High Pressures
This article presents a computational constitutive model for glass subjected to large
strains, high strain rates and high pressures. The model has similarities to a previously
developed model for brittle materials by Johnson, Holmquist and Beissel (JHB model),
but there are significant differences. This new glass model provides a material strength
that is dependent on the location and/or condition of the material. Provisions are made
for the strength to be dependent on whether it is in the interior, on the surface (different
surface finishes can be accommodated), adjacent to failed material, or if it is failed. The
intact and failed strengths are also dependent on the pressure and the strain rate. Thermal softening, damage softening, time-dependent softening, and the effect of the third
invariant are also included. The shear modulus can be constant or variable. The pressure-volume relationship includes permanent densification and bulking. Damage is accumulated based on plastic strain, pressure and strain rate. Simple (single-element)
examples are presented to illustrate the capabilities of the model. Computed results for
more complex ballistic impact configurations are also presented and compared to experimental data. [DOI: 10.1115/1.4004326]

Introduction

This article presents a computational constitutive model for


glass. Glass has been the focus of extensive research over the
years and much has been documented regarding its unique behavior. Bridgman [1,2] provides some of the earliest work where he
identifies three important characteristics of glass; it is very compressible, the bulk modulus decreases over a specified pressure
region, and a portion of the volumetric strain is unrecoverrable
(which produces permanent densification). Many researchers have
continued the investigation into the compressibility of glass with
some documenting up to 20% permanent densification [36].
There is also evidence that densification is dependent on temperature and shear [5,6]. Mackenzie [6] provides data that shows
increased densification when shear stress and/or temperature is
added. One of the more unique characteristics of glass is its very
high internal tensile strength. Rosenberg et al. [7] performed tensile-spall plate-impact experiments on soda lime glass and was
not able to produce spall and concluded that the spall strength
exceeded 5 GPa. Brar et al. [8] performed spall experiments on
soda lime and borosilicate glass and was unable to produce spall
in either one. Cagnoux [9,10] performed spall experiments on borosilicate glass and documented a high spall strength of 1.8 GPa on
average. Then there is the failure wave phenomenon that has
generated a considerable amount of research [11,12]. Although
much has been learned regarding this phenomenon, there are still
uncertainties regarding the correlation between opacity, damage
and stress state. Glass is also surface finish dependent, where a
smooth surface is stronger than a rough surface. Nie et al. [13]
demonstrated this effect by performing flexural tests on glass prepared with a very rough surface and with a surface etched using
hydrofluoric acid. The rough surface failed at a tensile stress of
approximately 100 MPa and the etched surface failed at 1200
MPa. Glass also exhibits scale effects where smaller samples are

Contributed by Applied Mechanics Division of ASME for publication in the JOURAPPLIED MECHANICS. Manuscript received October 15, 2010; final manuscript
received May 6, 2011; published online July 27, 2011. Assoc. Editor: Bo S. G. Janzon.

NAL OF

Journal of Applied Mechanics

stronger than larger samples. Anderson et al. [14] demonstrated


the effect of scale by performing ballistic impact experiments
onto thin glass plates. The smaller-scale glass plates were stronger, providing more resistance to penetration. Sun et al. [15] also
demonstrated the effect of scale by impacting small metal spheres
onto thin glass plates. The smaller-scale plates were less damaged
than the larger-scale plates. Wereszczak et al. [16] demonstrated
size effects by performing indentation tests onto soda lime glass.
The tensile strength increased significantly as the scale was
reduced. Wereszczak et al. concluded that this was the effect of
surface flaws and the statistical nature of their occurrence (larger
scales are more likely to have larger flaws and thus fail at a lower
stress). Glass, like other brittle materials, loses strength when
damaged. Chocron et al. [17] performed compression tests on
undamaged and predamaged borosilicate glass. The predamaged
samples were weaker than the undamaged samples. Chocron et al.
also demonstrated the significant effect that pressure has on the
strength, where higher pressures produce higher strengths. Glass
also appears to lose strength under one-dimensional plate-impact
loading [18,19], possibly a result of damage softening and/or thermal softening. There is also evidence that the failure process is
time dependent. Simha and Gupta [18] performed plate-impact
experiments on soda lime glass and concluded that the failure process was time dependent. Sunderam [19] performed pressureshear plate-impact experiments on soda lime glass and also
produced, what appeared to be, time-dependent failure. Glass also
exhibits a softening of the bulk modulus under compression. This
was first identified by Bridgeman [1] through quasi-static experiments and later confirmed through plate-impact tests [20]. There
is also evidence that the shear modulus softens. This is evident in
the Hugoniot Elastic Limit (HEL) for borosilicate glass [21] that
can only be reproduced using a reduced shear modulus. The
strength of glass is also dependent on the intermediate principal
stress (often referred to as the 3rd invariant effect). Handen et al.
[22] produced significantly higher strengths in pyrex glass when
the stress state was on the compressive meridian compared to the
tensile meridian. Glass also appears to be strain-rate sensitive. Nie
et al. [13] produced higher tensile strengths as the loading rate

C 2011 by ASME
Copyright V

SEPTEMBER 2011, Vol. 78 / 051003-1

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

included. The shear modulus can be constant or variable, and the


pressure-volume relationship includes permanent densification
and bulking.
2.1 Strength. The strength portion of the model is shown in
the upper portion of Fig. 1. The available strength (von Mises
equivalent stress), r, is dependent on the pressure, P, the dimen_ e_o (for e_o 1:0s1 ),
sionless equivalent (total) strain rate, e_ e=
the location of the material (interior, surface or adjacent to failed
material), and the damage, D. For undamaged material, D 0; for
partially damaged material, 0 < D < 1.0; and for fully damaged
(failed) material, D 1.0.
There are three curves that represent the strength of the intact
material. The reference intact strength represents the strength for
material adjacent to failed material, or effectively adjacent to a
surface with a very rough finish. This is the minimum intact
strength (when compared to the interior and surface intact
strengths). For a dimensionless strain rate of e_ 1:0, the reference intact strength goes (linearly) from r 0 at a tensile pressure of P  T, to a strength of r ri at a pressure of P Pi. T is
the maximum hydrostatic tension the reference strength can withstand. For pressures greater than Pi the strength is represented by
the analytic expression
r ri rmax  ri f1:0  expai P  Pi g

(1)

ai ri =rmax  ri Pi T 

(2)

where

Fig. 1

Description of the glass model

was increased. Glass is also sensitive to temperature. Chen [23]


demonstrated increased ductility and decreased strength in borosilicate glass when the temperature approached the glass transition
temperature, Tg. Glass also produces dilatancy, or bulking, during
failure. Glenn et al. [24] state that glass can bulk as much as 20%
or more under impact loading. Lastly, glass can produce dwell
and interface defeat as demonstrated by Behner et al. [25] and
Anderson et al. [26]. Borosilcate glass defeated a gold rod
(impacting at approximately 890 m/s) at the glass surface, producing no significant penetration into the glass target.
Glass is clearly a very complex material and modeling it
presents a significant challenge. The glass model presented herein
attempts to account for many of the aforementioned glass characteristics. The remainder of this article presents a description of the
glass model, a demonstration of various features of the model, and
computed results for several ballistic impact configurations that
are compared to experimental data.

This form provides a continuous slope at Pi and it limits the maximum strength to rmax (for e_ 1:0).
The strength for the surface and the strength for the interior
have similar forms and are defined by two dimensionless constants Dsurf and Dint respectively. Dsurf is a factor 0:01
 Dsurf  1:0 to allow the strength to be increased for material
on the surface that is not adjacent to failed material. This increase
becomes more pronounced as the pressure becomes more tensile.
For Dsurf 0.01 (the minimum allowed for numerical considerations) the surface strength is slightly below rmax , and for
Dsurf 1.0 the surface strength is identical to the reference intact
strength. For intermediate values of Dsurf a reference angle
href < 90o is determined that defines the angle between the horizontal line at rmax and the extrapolated line formed for the linear
relationship between the strength at r 0 (at P  T) to the
strength at r ri (at P Pi) as illustrated in Fig. 2. The corresponding angle for the strength defined by Dsurf is
hsurf Dsurf  href . The two extrapolated lines (defined by href

Description of the Model

A summary of the glass model is presented in Fig. 1. The model


has three components; one to describe the strength, one to
describe damage, and one to describe the pressure-volume behavior. Although the model is similar to a previously developed
model for brittle materials by Johnson, Holmquist and Beissel
(JHB model) [27], there are many differences. This new glass
model has an intact strength that is dependent on whether the material is located in the interior, on the surface, or adjacent to failed
material. There is also a failed strength which represents the
strength of the material that is fully damaged (D 1.0). Thermal
softening, damage softening (instantaneous or gradual), time-dependent softening, and the effect of the third invariant are also
051003-2 / Vol. 78, SEPTEMBER 2011

Fig. 2 Description of the interior, surface and reference


strength

Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

and hsurf ) meet at a common point at the strength of rmax . This


at Pi. For pressures greater than Pi the analytic
also defines rsurf
i
expressions presented in Eqs. (1) and (2) are used but ri is
and T is replaced by Tsurf tan1 hsurf
replaced by rsurf
i

P
.
For
pressures
less than Pi the interior strength
 rsurf
i
i
to Tsurf.
decreases linearly from rsurf
i
The strength for the interior is defined in a similar manner. Dint
is a factor 0:01  Dint  1:0 to allow the strength to be
increased for material in the interior that is not adjacent to failed
material. Again, this increase becomes more pronounced as the
pressure becomes more tensile. For Dint 0.01 the interior
strength is slightly below rmax , and for Dint 1.0 the interior
strength is identical to the reference intact strength. For intermediate values of Dint a reference angle href < 90o is determined, as
defined previously. The corresponding angle for the strength
defined by Dint is hint Dint  href . This also defines rint
i at Pi. For
pressures greater than Pi the analytic expressions presented in
and T is
Eqs. (1) and (2) are used but ri is replaced by rint
i
replaced by Tint tan1 hint  rint
i  Pi . For pressures less than
Pi the interior strength decreases linearly from rint
i to Tint.
The strength for the failed material has a similar form to that of
the reference intact strength. It goes linearly from r 0 at P 0,
to r rf at P Pf. For pressures greater than Pf the strength is
represented by




 
(3)
r rf rfmax  rf 1:0  exp af P  Pf

1
J3 s3x s3y s3z 3sx s2xy 3sx s2xz 3sy s2xy 3sy s2yz
3
3sz s2xz 3sz s2yz 6sxy syz sxz

(8)

The reduced intact strength is given by


rreduced
i
rinput
i


J3fact 
1 sin3h 1=J3fact 1  sin3h
2

(9)

is the input intact strength along the compressive


where rinput
i
meridian.
Some comments should be made regarding the algorithms used
to determine surface elements, interior elements and elements adjacent to failed material. If an element has a least one node on the
surface it is considered a surface element and Dsurf is used to
describe the intact strength. If an element has at least one node in
common with a failed element it is considered adjacent to failed
material and the reference intact strength is used to describe the
strength. If an element is not a surface element or an element adjacent to failed material it is an interior element and Dint is used to
describe the intact strength. Also, the factors Dsurf and Dint are not
used for particles (because it is not a straightforward procedure to
determine if a particle is on the surface, in the interior, or adjacent
to failed material). This is generally not a problem when conversion of elements into particles is used, because the material usually fails before it is converted.

where
af rf

. 

Pf rfmax  rf

(4)

The intact and failed strengths in Eqs. (1)(4) are for a dimensionless strain rate of e_ 1:0. If ro is the strength at e_ 1:0, then
the strength at other strain rates is
r ro 1:0 C ln e_

(5)

where C is the dimensionless strain rate constant.


Thermal softening is included through the thermal softening
exponent M. If M > 0 the material is softened by the factor

(1  T M) for both the intact and failed material. The homologous

temperature is T (Tc  Tr)/(Tg  Tr) where Tc is the current
computed temperature, Tr is the room temperature and Tg is the
glass transition temperature. This form is used because it is
straightforward to incorporate into the strength model, requires
only one constant, M, and provides good agreement with the test
data from Chen [23].
Lastly, the effect of the third invariant is described. The implementation is similar to that presented by Frank and Adley [28]. It
applies to both the intact and failed strengths. When the third
invariant is used, the intact strength should be defined for a stress
state in the compressive meridian (r3 < r1 and r1 r2 ) where
r1 , r2 and r3 are the principal stresses (tension is positive). The
strength for the tensile meridian is reduced by a dimensionless
constant J3fact times the strength in the compressive meridian
where 0.5 < J3fact < 1.0. The stress state for the tensile meridian is
(r1 > r2 and r2 r3 ).
The interpolation between the compressive and tensile meridians is provided by the Lode angle (h) which can be defined as a
function of the second and third invariants (J2 and J3)
p
3 3 J3

(6)
sin3h
2 J 3=2
2

where J2 and J3 are defined in terms of the deviator stresses


1
J2 s2x s2y s2z 2s2xy 2s2yz 2s2xz
2
Journal of Applied Mechanics

(7)

2.2 Damage. The damage for failure is shown in the lower


left portion of Fig. 1 and it is accumulated as follows:

(10)
D R Dep =e fp
where Dep is the increment of equivalent plastic strain during the
current cycle of integration and e fp is the plastic strain to failure
under the current dimensionless pressure, P P=rmax . The general expression for the failure strain is


e fp D1 P T  N 1 Cf ln e_

(11)

where D1, N and Cf are dimensionless constants and T* T/rmax.


The strain rate constant Cf provides an increase in the strain to
failure for elevated strain rates. Tensile hydrostatic pressures
greater than T will not cause the material to fail (D 1.0) instantaneously as is the case for the JHB model [27]. Here, very large
tensile pressures are possible if the material remains elastic, but
any amount of plastic strain will produce failure. This allows the
strength of the interior and surface strengths to be reached at tensile pressures greater than T. This damage formulation assumes
that damage does not accumulate when the material is in an elastic
stress state. The authors are not aware of any quantitative data
which indicates a weakening of the material when it is within the
yield surface.
When D 1.0 the material fails and the strength is reduced as
shown in the top portion of Fig. 1. Two types of softening (damage softening or time-dependent softening) are provided to allow
for variations in how the material transitions from the intact surface to the failed surface. Damage softening allows the material to
soften gradually, rather than fail instantaneously at D 1.0. Material begins to soften when D (1.0 Dsoft) and then linearly softens to the failed strength at D 1.0 where Dsoft is a dimensionless
coefficient (0 < Dsoft < 1.0). In the softening region [D > (1.0
Dsoft)] the bulking pressure is added incrementally, in a manner
similar to that used for the JH-2 model for brittle materials [29].
For Dsoft 0 there is no gradual softening and the material fails
instantaneously at D 1.0. Alternatively, a time-dependent softening can be used. This option gradually reduces the strength over
SEPTEMBER 2011, Vol. 78 / 051003-3

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

DU Ui  Uf

a specified time, tfail. The material begins to soften in a linear


manner when D 1.0.
2.3 Pressure. Glass exhibits permanent densification when
subjected to high pressures [26] and bulking when failure occurs
[24]. The hydrostatic pressure, with permanent densification and
bulking, is shown in the lower right portion of Fig. 1. Permanent
densification begins when l > lelastic and is complete when
l llock. The pressure for loading only (no unloading), no bulking (b 0, defined later) and before complete densification
(l < llock), is
P K1 l K2 l2 K3 l3

(12)

where K1, K2 and K3 are constants (K1 is the bulk modulus); and
l Vo =V  1 q=qo  1 for current volume and density (V and
q), and initial volume and density (Vo and qo ). The pressure for
l > llock is simply
P Plock K4 l  llock

(13)

where Plock is the pressure at llock, and K4 and llock are constants.
The process of unloading is more complex and is a function of
the magnitude of the permanent densification, lperm, and the maximum volumetric strain attained, lmax. When lmax < lelastic there
is no permanent densification and unloading occurs along the
loading path described by Eq. (12). When lmax  llock permanent
densification is complete and unloading occurs along the permanent densification curve (from llock to lperm). The unloading path
(from llock to lperm) is determined by interpolating between the
response provided by Eq. (12) and the response provided by
Eq. (13) (when extrapolated down to lolock ). The interpolation is
based on the ratio lperm =lolock and produces unloading along a path
described by Eq. (12) when lperm =lolock 0, and by Eq. (13) when
lperm =lolock 1.0.
There is also the possibility to unload in the transition zone
(lelastic < lmax < llock ). The unloading path is interpolated
between the path describe by Eq. (12) and the permanent densification curve previously described. The interpolation is based on
the ratio / lmax  lelastic =llock  lelascic and produces
unloading along a path described by Eq. (12) when / 0, and
along the permanent densification curve when / 1.0.
For tensile pressures 
l < 0, P K1 l where K 1 K1 and
l l when lmax < lelastic, K1 is the stiffness at lperm and
l l  lperm when lmax  llock , and in the transition zone
(lelastic < lmax < llock ) K1 is interpolated and l l  l0 where
l0 lperm lmax  lelastic =llock  lelascic is the permanent volumetric strain at P 0.
During the softening/failure process (from damage or time-dependent softening) bulking (pressure and/or volume increase) can
occur. This requires an additional pressure increment, DP, such
that
P K1 l K2 l2 K3 l3 DP

(14)

The pressure increment is determined from energy considerations.


Looking back to the strength model in Fig. 1, there is a decrease
in strength when the material goes from an intact state (D < 1.0)
to a failed state (D 1.0) (for instantaneous failure with Dsoft 0).
This represents a decrease in the elastic internal energy of the deviator stresses. The general expression for this elastic internal
energy (of the deviator stresses) is
U r2 =6G

(15)

where r is the von Mises equivalent stress and G is the elastic


shear modulus. The corresponding decrease in internal energy can
be expressed as
051003-4 / Vol. 78, SEPTEMBER 2011

(16)

where Ui is the internal energy of the intact material before failure


(D < 1.0) and Uf is the internal energy after failure (D 1.0).
This internal energy loss (of the deviator stresses) can be converted to potential hydrostatic internal energy by adding DP. An
approximate equation for the internal energy conservation is
 bDU
DPlf DP2 =2K

(17)

where lf is the current value of l at failure, K is the effective bulk


modulus, and b is the fraction (0  b  1:0 of the internal (deviator) energy loss converted to potential hydrostatic energy. The
first term (DPlf ) is the approximate potential energy for l > 0 and
the second term [DP2 =2K] is the corresponding potential energy
for l < 0.
Solving for DP gives
q



 f 2 2bKDU
 f
(18)
Kl
DP Kl
The bulking pressure is computed only for failure under compression (lf > 0). Note that DP 0 for b 0 and that DP increases
as DU increases and/or lf decreases. When there is gradual damage softening (Dsoft > 0), DP is updated during the course of the
softening process [29].
In addition to a constant shear modulus (which implies a variable Poissons ratio), there is a provision to use a variable shear
modulus, Gvar, that is proportional to the current bulk modulus,
Kcur and is given by


3 1  2
(19)
Gvar Kcur
2 1
where  is Poissons ratio. This relationship is based on the
assumption of a constant Poissons ratio.
It should be noted that the pressure does not include energy dependence (other than bulking) such as occurs in the Mie-Gruneisen and other Equations of State. Generally, the energy effects
are more important for very high pressures, and are not significant
for ballistics problems which involve lower pressures. The inclusion of energy effects would be very complex when combined
with permanent densification and bulking, and it is not clear how
these three effects (internal energy, permanent densification, bulking) could be combined. Furthermore, the authors are not aware of
any test data from which the appropriate constants could be determined. The bulking algorithm has been used with the JHB model
for ceramics [27] and it provides a qualitative description of
behavior during and after failure. The material rapidly expands
when failed under low confinement/pressure, and it develops a
distinct pressure increase when it is confined. Similar behavior
has been observed when glass fails. Two desirable features of this
bulking algorithm are that it does not require any additional constants (other than the fraction of energy that goes into bulking),
and that internal energy is conserved. For penetration problems at
ballistic velocities, the presence of bulking does not appear to
have a significant effect on the penetration velocities and depths.
The effect of strength is generally much more important.

Demonstration of the Model

Several features of the model are illustrated by the examples


shown in Figs. 36. The examples presented in Figs. 35 demonstrate the effect of densification, damage softening and the intermediate principal stress (3rd invariant) respectively, and are
produced by axially loading (and unloading) a single element in
uniaxial strain. The example presented in Fig. 6 demonstrates the
effect of time-dependent failure and high-internal-tensile strength
using computed results from spall plate-impact computations.
Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

erosion occurs. It is not clear why the glass appears stronger in


this velocity regime and then suddenly appears much weaker at
Vs > 600 m/s. Although the computed results do not capture this
response, it may be possible to produce better agreement using a
different set of constants (specifically T, tfail, Dint and Dsurf). General comparisons can also be made regarding the computed and
experimental post-mortem targets. The tested targets consistently
produced rear surface spall cones and a significant amount of radial and circumferential cracks emanating from the impact location (although the cracking was significant, the targets remained
intact). The computed targets also produced rear surface spall
cones, but the cracking was much less than what was observed
experimentally (this could be due to the computations being performed in 2D axisymmetry where radial cracking cannot occur).
Figure 11 demonstrates the capability to compute size effects.
Computed results are presented for the 0.50-cal (scale 1.0) and
the 0.22-cal (scale 0.44) projectile for Vs 300 m/s and 400
m/s. At 300 m/s, the 0.50-cal projectile exits the target at
Vr 22 m/s but the smaller scale 0.22-cal projectile is stopped
(there is also significantly more projectile erosion). As the
impact velocity is increased the computed size effect diminishes
as shown at Vs 400 m/s. The ability for the computations to
produce size effects is due to the time-dependent features in the
glass model. Smaller scale events happen in a shorter time period resulting in larger strain rates for the smaller scale. The
strain rate effect in the ductility (Cf), the strain rate effect in the
strength (C), and the finite time-to-fail (tfail), all contribute to
making smaller scales stronger.

Summary and Conclusions

This article has presented a new computational constitutive


model for glass subjected to large strains, high strain rates and
high pressures. This new glass model has an intact strength and a
failed strength that are dependent on the pressure and the strain
rate, damage softening or time-dependent softening that transitions the strength from intact to failed, thermal softening, a constant or variable shear modulus, the effect of the third invariant, a
pressure-volume relationship that includes permanent densification and bulking, and damage that is accumulated from plastic
strain, pressure and strain rate. The intact strength is also dependent on the location of the material (whether it is in the interior, on
the surface or adjacent to failed material). Significant features of
the model are the ability to compute size effects (smaller is stronger), surface effects, and high internal tensile strength.
Several features of the model were illustrated using a simple
(single-element) problem. Computations were also presented for
more complex, high-velocity-impact conditions using preliminary
constants for borosilicate glass determined from the literature.
The computed results demonstrate the ability of the model to produce reaseonable results for a broad range of impact conditions.

Acknowledgment
This work was performed for the U. S. Army RDECOMTARDEC under Contract No. W56HZV-06-C-0194. The authors
would like to thank D. Templeton and R. Rickert (U. S. Army
TARDEC) and to C. Gerlach (Southwest Research Institute) for
their contributions to this work. The authors would like to especially thank C. Anderson, Jr. for his helpful discussions and experimental work in support of this effort.

References
[1] Bridgman, P. W., 1948, The Compression of 39 Substances to 100,000 kg/
cm2, Proceedings of the American Academy of Arts and Sciences, 76(3), pp.
5587.
[2] Bridgman, P. W. and Simon, I., 1953, Effects of Very High Pressures on
Glass, J. Appl. Phys., 24(4), pp. 405413.

Journal of Applied Mechanics

[3] Uhlmann, D. R., 1973, Densification of Alkali Silicate Glasses at High Pressure, J. Non-Cryst. Solids, 13, pp. 8999.
[4] Rouxel, T., Ji, H., Hammouda, T., and Moreac, A., 2008, Poissons Ratio and
the Densification of Glass Under High Pressure, Phys. Rev. Lett., 100,
p. 225501.
[5] Sakka, S. and Mackenzie, J. D., 1969, High Pressure Effects on Glass, J.
Non-Cryst. Solids, 1, pp. 107142.
[6] Mackenzie, J. D., 1963, High-Pressure Effects on Oxide Glasses: I, Densification in Rigid State, J. Am. Ceram. Soc., 46(10), pp. 461470.
[7] Rosenberg, Z., Yaziv, D., and Bless, S., 1985, Spall Strength of Shock-Loaded
Glass, J. Appl. Phys., 58(8), pp 32493251.
[8] Brar, N., Rosenberg, S., and Bless, S., 1991, Spall Strength and Failure Waves
in Glass, J. Phys. (Paris), 1, pp. 639634.
[9] Cagnoux, J., 1985, Deformation et Ruine dun Verre Pyrex Soumis a un Choc
Intense: Etude Experimentale et Modelisation du Comportement, Ph.D. thesis,
LUniversite de Poitiers.
[10] Cagnoux, J. and Longy, F., 1988, Spallation and Shock-Wave Behaviour of
Some Ceramics, J. Phys. (Paris), 49(9), pp. 310.
[11] Kanel, G. I., Bogatch, A. A., Razorenov, S. V., and Chen, Z., 2002,
Transformation of Shock Compression Pulses in Glass due to the Failure
Wave Phenomena, J. Appl. Phys., 92(9), pp. 50455052.
[12] Anderson, C. E., Jr., Orphal, D. L., Behner, T., and Templeton, D. W., 2009,
Failure and Penetration Response of Borosilicate Glass During Short-Rod
Impact, International Journal of Impact Engineering, 36, pp. 789798.
[13] Nie, X., Chen, W., Wereszczak, A., and Templeton, D., 2009, Effect of Loading Rate and Surface Conditions on the Flexural Strength of Borosilicate
Glass, J. Am. Ceram. Soc., 92(6), pp. 12871295.
[14] Anderson, C. Jr., Weiss, C., and Chocron, S., 2009, Impact Experiments into
Borosilicate Glass at Three Scale Sizes, Southwest Research Institute, San
Antonio, TX, Technical Report No. 18.12544/018.
[15] Sun, X., Khaleel, M., and Davies, R., 2005, Modeling of Stone-Impact Resistance of Monolithic Glass Ply using Continuum Damage Mechanics, Int. J.
Damage Mech., 14, pp. 165178.
[16] Wereszczak, A. A., Kirkland, T. P., Ragan, M. E., Strong, K. T., Jr., and Lin,
H., 2010, Size Scaling of Tensile Failure Stress in a Float Soda-Lime-Silicate
Glass, International Journal of Applied Glass Science, 1(2), pp. 143150.
[17] Chocron, S., Anderson, C. E., Jr., Nicholls, E., and Dannemann, K. A.,
Characterization of Confined Intact and Damaged Borosilicate Glass, J. Am.
Ceram. Soc. (to be published).
[18] Simha, C. and Gupta, Y., 2004, Time-Dependent Inelastic Deformation of
Shocked Soda-Lime Glass, J. Appl. Phys., 96(4), pp. 18801890.
[19] Sundaram, S., 1993, Pressure-Shear Plate Impact Studies of Alumina Ceramics
and the Influence of an Intergranular Glassy Phase, Ph.D. thesis, Brown University, Providence, RI.
[20] Cagnoux, J., 1982, Shock-Wave Compression of a Borosilicate Glass up to
170 kbar, Shock Compression of Condensed Matter-1982, pp. 392296.
[21] Alexander, C. S., Chhabildas, L. C., Reinhart, W. D., and Templeton, D. W.,
2008, Changes to the Shock Response of Fused Quartz due to Glass Modification, International Journal of Impact Engineering, 35, pp. 13761385.
[22] Handin, J., Heard, H. C., and Magouirk, J. N., 1967, Effects of the Intermediate Principal Stress on the Failure of Limestone, Dolomite, and Glass at Different Temperatures and Strain Rates, J. Geophys. Res., 72(2), pp. 611640.
[23] Chen, W., 2010, Purdue University, private communication.
[24] Glenn, L. A., Moran, B., and Kusubov, A. S., 1990, Modeling Jet Penetration
in Glass, Lawrence Livermore National Laboratory, CA., Technical Report
No. UCRL-JC-103512.
[25] Behner, T., Anderson, C., Jr., Orphal, D., Hohler, V., Moll, M., and Templeton,
D., 2008, Penetration and Failure of Lead and Borosilicate Glass against Rod
Impact, International Journal of Impact Engineering, 35, pp. 447456.
[26] Anderson, C. E., Jr., Behner, Th., Holmquist, T. J., Wickert, M., Hohler, V.,
and Templeton, D. W., 2007, Interface Defeat of Long Rods Impacting Borosilicate Glass, Proceedings of the 23rd International Symposium on Ballistics,
Tarragona, Spain, pp. 10491056.
[27] Johnson, G. R., Holmquist, T. J., and Beissel, S. R., 2003, Response of Aluminum Nitride (Including a Phase Change) to Large Strains, High Strain Rates,
and High Pressures, J. Appl. Phys., 94(3), pp. 16391646.
[28] Frank, A. and Adley, M., 2007, On the Importance of a Three-Invariant Model
for Simulating the Perforation of Concrete Targets, Proceedings from the 78th
Shock and Vibration Symposium, Philadelphia, PA.
[29] Johnson, G. R. and Holmquist, T. J., 1994, An Improved Computational Constitutive Model for Brittle Materials, High Pressure Science and Technology
1993, S. C. Schmidt, J. W. Schaner, G. A. Samara, and M. Ross, eds., AIP,
New York, pp. 981984.
[30] Johnson, G. R., Stryk, R. A., Holmquist, T. J., and Beissel, S. R., 1997,
Numerical Algorithms in a Lagrangian Hydrocode, Wright Laboratory, FL,
Technical Report No. WL-TR-1997-7039.
[31] Johnson, G. R., Beissel, S. R., and Stryk, R. A., 2002, An Improved Generalized Particle Algorithm that Includes Boundaries and Interfaces, Int. J. Numer.
Methods Eng., 53, pp. 875904.
[32] Johnson, G. R., Stryk, R. A., Beissel, S. R., and Holmquist, T. J., 2002,
Conversion of Finite Elements into Meshless Particles for Penetration Computations Involving Ceramic Targets, Shock Compression of Condensed
Matter2001, M. D. Furnish, N. N. Thadhani, and Y. Horie, eds., AIP, New
York.

SEPTEMBER 2011, Vol. 78 / 051003-9

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

intact surface is again encountered. Unloading continues plastically from points 6 to 7. At point 7, both the axial deviator stress
and the pressure go to zero. For J3fact 0.5 the unloading is different because a lower intact surface is used for the tensile meridian. Elastic unloading continues from points 4 to 5 where the
lower intact strength is encountered. Unloading continues plastically from Points 5 to 7. At point 7, both the axial deviator stress
and the pressure go to zero.
Figure 6 presents the effect of time-dependent failure and the
ability to develop high-internal-tensile strength. Four tensile-spall
plate-impact computations are used to demonstrate these effects.
The material constants are the same as those presented previously
except the interior strength is increased to Dint 0.7 (this provides
a significant increase in the interior tensile strength and corresponds to a spall strength of approximately 450 MPa). The plate
impact configuration consists of a 5 mm glass impactor and a 15
mm glass target (this configuration will produce a tensile stress 5
mm from the rear target surface). The impact velocity is 200 m/s
and the particle velocity-time histories are taken at the rear target
surface. In the upper left quadrant of Fig. 6 is the computed wave
profile for tfail 0.4 ls, in the lower left is the wave profile for
tfail 0.2 ls, in the upper right is the wave profile for tfail 0.02
ls and in the lower right is the wave profile for tfail 0.
Before the results in Fig. 6 are discussed, a brief overview of a
tensile-spall plate-impact test is provided. In a spall experiment,
an impactor strikes a target producing an elastic compressive
wave that travels into the target and into the impactor. The elastic
compressive wave reflects off the rear surface of the target and
propagates back toward the impact surface. Due to wave interactions a tensile stress is created in the interior of the target located
one impactor thickness from the rear target surface (this is true
when both the impactor and target are made of the same material).
If the material can withstand the magnitude of the tensile stress,
the material does not fail and no spall is produced. If the material
cannot withstand the tensile stress, the material fails, creating a
pullback signal which travels back to the rear surface where it is
measured. The magnitude of the pullback signal is proportional to
the tensile stress at which the material fails (the spall strength,
rspall). Since spall occurs inside the material (but is measured at
the rear surface), a spall experiment provides an indirect measure
of the internal tensile strength.
Returning to Fig. 6, the computed result for tfail 0.4 ls is discussed first. This response is described with the help of the stresspressure insert in the lower right quadrant in Fig. 6 (the stresspressure response is taken at the spall location, 5 mm from the
rear surface). The impactor strikes the target at 200 m/s and the
stress goes from point 1 to point 2. The peak compressive stress is
elastic and is maintained until the arrival of the release wave. The
release wave rapidly takes the material into tension (at the spall
location). The stress goes from point 2 to point 1 to point 3 where
the interior intact strength is encountered. The material can develop no plastic strain and failure occurs producing the pullback
signal. The magnitude of the pullback signal is proportional to the
spall stress of 450 MPa. Since the tensile stress occurs in the interior of the target the interior intact strength governs the response
(point 3) and not the reference intact strength (point 4). Also note
the time delay of approximately 0.4 ls in the pullback signal, due
to the time-dependent failure. The result for tfail 0.2 ls is similar
to that of tfail 0.4 ls except the duration of the pullback signal is
reduced, which is due to the reduction in the time it takes the material to fail. The result for tfail 0.02 ls is more complex. The
magnitude of the pullback signal is significantly smaller than produced for tfail 0.4 ls and tfail 0.2 ls. This is due to the very
short duration of failure, the condition of the material and the
propagation of the pullback signal. When failure occurs at the
spall plane a new surface is created. This new surface is adjacent
to failed material (created at the spall plane) and is governed by
the reference intact strength. If the time to failure is sufficiently
short, the magnitude of the pullback signal cannot be maintained
as it propagates to the rear surface. The pullback signal will
051003-6 / Vol. 78, SEPTEMBER 2011

continue to decay until either the reference intact strength is


reached or the signal arrives at the rear surface. The final example
is for tfail 0 and produces only a small pullback signal. Even
though the spall strength of 450 MPa is attained at the spall plane,
it cannot be maintained and quickly decays to the stress represented by the reference intact strength at point 4.

4 Example Computations for High-Velocity Impact


Conditions
The following computed results are for more complex, highvelocity-impact conditions. The computed results use the new
glass model and preliminary constants for borosilicate glass estimated from the literature. The determination of constants is not a
straightforward process because some of the constants cannot be
determined explicitly from the test data. The general approach to
determine constants for the glass model is as follows: The pressure-volume response, the maximum intact strength, the time
required for the material to fail, and the shear modulus (constant
or variable) are estimated from compression plate-impact experiments where both the longitudinal and lateral stresses are
measured. The interior strength is estimated from tensile-spall
plate-impact experiments. Quasi-static and dynamic four-point
bending flexural tests are used to estimate the reference strength,
the surface strength and the strain-rate effect for damage. Quasistatic and dynamic compression tests are used to estimate the
strain-rate effect on the strength. Confined quasi-static compression tests and pressure-shear plate-impact tests are used to estimate the strength of the failed material. The damage model
constants are estimated by providing computed results that are in
good agreement with ballistic experiments that produce dwell,
interface defeat and penetration. Finally, although it is clear that
glass undergoes permanent densification [36], a procedure to
determine the onset, completion and magnitude of permanent densification has not been identified. This will be an area of future
research.
The objective of this section is not so much to provide a measure of the accuracy of the model but rather to demonstrate its
ability to produce reasonable results for a broad range of impact
conditions using one set of constants. All the computations were
performed with the 2010 version of the EPIC code using onedimensional (1D) and two-dimensional (2D) Lagrangian finite
elements [30] and 2D meshless particles [31]. For the 2D example
computations involving severe distortions, the finite elements are
automatically converted into meshless particles during the course
of the computation [32].
4.1 Plate Impact. Figure 7 shows a comparison of the computed results and experimental results for two compressive plateimpact tests (BORO-9 is offset by 1.0 ls for clarity). The compression tests (BORO-9 and BORO-10) are provided by
Alexander et al. [21] and use a copper impactor and a borosilicate
target backed by lithium fluoride. The particle velocity is presented as a function of time recorded at the borosilicate-lithium
fluoride interface. For both compressive tests there is reasonable
agreement between the computed and experimental results. A few
comments should be noted regarding the computed wave profiles;
the ramping of the initial elastic wave is due to a softening of the
bulk modulus, the hugoniot elastic limit (HEL) has a magnitude
of approximately up 0.6 km/s, and the increases in the particle
velocity (at approximately 3 ls for BORO-10 and at 4 ls for
BORO-9) are a result of wave reflections at the impactor-target
interface.
4.2 Interface Defeat and High Velocity Penetration. Figure 8
presents computed results for a gold projectile impacting a borosilicate target with a copper buffer attached to the impact surface.
The buffer is included to attenuate the impact shock and produce
gradual loading on the glass. The computed results produce
Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

Fig. 7 Comparison of the computed results and experimental


results for two plate-impact tests

interface defeat (no glass penetration) at Vs 800 m/s and prompt


penetration at Vs 900 m/s. This is consistent with the test results
provided by Anderson et al. [26].
Behner et al. [25] performed experiments using no buffer (gold
rod impacting bare glass) producing penetration at much lower
impact velocities. For these experiments the penetration velocity
was determined using a series of flash x-rays and are presented as
a function of impact velocity as shown in the lower portion of
Fig. 9. Also shown are the computed penetration velocities for
Vs 900 m/s and Vs 2400 m/s. The computed penetration velocity is slightly low compared to the test data at Vs 900 m/s, but is
in closer agreement at Vs 2400 m/s. In the top portion of Fig. 9,
the computed results show material damage at 30 ls (for Vs 900
m/s) and at 10 ls (for Vs 2400 m/s). Behner et al. [25] also provides flash x-ray images and high-speed photographs at similar
times and impact velocities as those presented in the top portion
of Fig. 9. A comparison of the computed responses to the x-ray
and optical images yield the following observations: The computed and experimental cavity sizes were generally in good agreement for both impact velocities (as defined by the flow of the high-

Fig. 8 The initial geometry and computed results for a gold


rod impacting a borosilicate target with a copper buffer at
V 5 800 m/s and V 5 900 m/s

Journal of Applied Mechanics

Fig. 9 Comparison of computed and experimental results for a


gold rod impacting bare borosilicate glass

density projectile debris). The lower impact velocity produced a


larger and less uniform cavity, the higher impact velocity produced
a smaller more uniform cavity. Both the computed and experimental results produced a significant amount of glass fragments moving
out from the impact surface (although there appeared to be more
fragments produced in the experiments than in the computations).
Lastly, the outer edge of the experimental targets expand radially
outward during penetration, but the computed results did not (this
could be due to the computations being performed in 2D axisymmetry where radial cracking cannot occur).
4.3 Steel Projectiles Impacting Thin Glass Targets. Anderson
et al. [14] performed experiments using a pointed steel projectile
impacting thin plates of borosilicate glass at three different scale
sizes. For the largest scale (scale 1.0) the projectile was 12.7
mm in diameter and 38.1 mm long (0.50-cal) and the target was
21.0 mm thick and 457.2 mm square. For the smaller scales
(scale 0.75 and 0.44) both the projectile and target plates were
scaled accordingly. The experimental results produced some significant findings: The targets appeared stronger as the scale was
reduced (particularly at the lower impact velocities); glass is very
strong producing significant projectile erosion; and there appears
to be a change in the glass behavior at approximately Vs 500
600 m/s, which produced a discontinuity in the Vs-Vr curve (this is
also where the maximum projectile erosion occurs). The computations focused primarily on the 0.50-cal experiments (scale 1.0),
although the 0.22-cal (scale 0.44) was also used to investigate
the scale effect.
Figures 10 and 11 present computed and experimental results
for steel projectiles impacting thin glass targets. Figure 10 shows
the computed results (showing material damage) for three impact
SEPTEMBER 2011, Vol. 78 / 051003-7

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

Fig. 10 Comparison of computed and experimental results for a steel projectile impacting borosilicate glass

velocities using the 0.50-cal projectile. At Vs 300 m/s the projectile exits the target with a very low velocity of 22 m/s and there
is a large spall cone produced on the rear target surface. There
also has been some erosion of the projectile. At Vs 400 m/s and
500 m/s the projectile exits the target at a much higher exit velocity and with less projectile erosion. The targets appear similar,
exhibiting large cone cracks that propagate from the target mid-

section to near the rear surface. The lower right portion of Fig. 10
presents a comparison of the residual velocity as a function of
impact velocity for the computed and experimental results. The
computed exit velocities are in good agreement at Vs 400 m/s
and for Vs > 600 m/s, but the targets are too strong for Vs < 400
m/s. The computed results also do not capture the target response
for 500 m/s  Vs  600 m/s where the maximum projectile

Fig. 11 Comparison of computed results for a steel projectile impacting borosilicate glass at
two scales

051003-8 / Vol. 78, SEPTEMBER 2011

Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

erosion occurs. It is not clear why the glass appears stronger in


this velocity regime and then suddenly appears much weaker at
Vs > 600 m/s. Although the computed results do not capture this
response, it may be possible to produce better agreement using a
different set of constants (specifically T, tfail, Dint and Dsurf). General comparisons can also be made regarding the computed and
experimental post-mortem targets. The tested targets consistently
produced rear surface spall cones and a significant amount of radial and circumferential cracks emanating from the impact location (although the cracking was significant, the targets remained
intact). The computed targets also produced rear surface spall
cones, but the cracking was much less than what was observed
experimentally (this could be due to the computations being performed in 2D axisymmetry where radial cracking cannot occur).
Figure 11 demonstrates the capability to compute size effects.
Computed results are presented for the 0.50-cal (scale 1.0) and
the 0.22-cal (scale 0.44) projectile for Vs 300 m/s and 400
m/s. At 300 m/s, the 0.50-cal projectile exits the target at
Vr 22 m/s but the smaller scale 0.22-cal projectile is stopped
(there is also significantly more projectile erosion). As the
impact velocity is increased the computed size effect diminishes
as shown at Vs 400 m/s. The ability for the computations to
produce size effects is due to the time-dependent features in the
glass model. Smaller scale events happen in a shorter time period resulting in larger strain rates for the smaller scale. The
strain rate effect in the ductility (Cf), the strain rate effect in the
strength (C), and the finite time-to-fail (tfail), all contribute to
making smaller scales stronger.

Summary and Conclusions

This article has presented a new computational constitutive


model for glass subjected to large strains, high strain rates and
high pressures. This new glass model has an intact strength and a
failed strength that are dependent on the pressure and the strain
rate, damage softening or time-dependent softening that transitions the strength from intact to failed, thermal softening, a constant or variable shear modulus, the effect of the third invariant, a
pressure-volume relationship that includes permanent densification and bulking, and damage that is accumulated from plastic
strain, pressure and strain rate. The intact strength is also dependent on the location of the material (whether it is in the interior, on
the surface or adjacent to failed material). Significant features of
the model are the ability to compute size effects (smaller is stronger), surface effects, and high internal tensile strength.
Several features of the model were illustrated using a simple
(single-element) problem. Computations were also presented for
more complex, high-velocity-impact conditions using preliminary
constants for borosilicate glass determined from the literature.
The computed results demonstrate the ability of the model to produce reaseonable results for a broad range of impact conditions.

Acknowledgment
This work was performed for the U. S. Army RDECOMTARDEC under Contract No. W56HZV-06-C-0194. The authors
would like to thank D. Templeton and R. Rickert (U. S. Army
TARDEC) and to C. Gerlach (Southwest Research Institute) for
their contributions to this work. The authors would like to especially thank C. Anderson, Jr. for his helpful discussions and experimental work in support of this effort.

References
[1] Bridgman, P. W., 1948, The Compression of 39 Substances to 100,000 kg/
cm2, Proceedings of the American Academy of Arts and Sciences, 76(3), pp.
5587.
[2] Bridgman, P. W. and Simon, I., 1953, Effects of Very High Pressures on
Glass, J. Appl. Phys., 24(4), pp. 405413.

Journal of Applied Mechanics

[3] Uhlmann, D. R., 1973, Densification of Alkali Silicate Glasses at High Pressure, J. Non-Cryst. Solids, 13, pp. 8999.
[4] Rouxel, T., Ji, H., Hammouda, T., and Moreac, A., 2008, Poissons Ratio and
the Densification of Glass Under High Pressure, Phys. Rev. Lett., 100,
p. 225501.
[5] Sakka, S. and Mackenzie, J. D., 1969, High Pressure Effects on Glass, J.
Non-Cryst. Solids, 1, pp. 107142.
[6] Mackenzie, J. D., 1963, High-Pressure Effects on Oxide Glasses: I, Densification in Rigid State, J. Am. Ceram. Soc., 46(10), pp. 461470.
[7] Rosenberg, Z., Yaziv, D., and Bless, S., 1985, Spall Strength of Shock-Loaded
Glass, J. Appl. Phys., 58(8), pp 32493251.
[8] Brar, N., Rosenberg, S., and Bless, S., 1991, Spall Strength and Failure Waves
in Glass, J. Phys. (Paris), 1, pp. 639634.
[9] Cagnoux, J., 1985, Deformation et Ruine dun Verre Pyrex Soumis a un Choc
Intense: Etude Experimentale et Modelisation du Comportement, Ph.D. thesis,
LUniversite de Poitiers.
[10] Cagnoux, J. and Longy, F., 1988, Spallation and Shock-Wave Behaviour of
Some Ceramics, J. Phys. (Paris), 49(9), pp. 310.
[11] Kanel, G. I., Bogatch, A. A., Razorenov, S. V., and Chen, Z., 2002,
Transformation of Shock Compression Pulses in Glass due to the Failure
Wave Phenomena, J. Appl. Phys., 92(9), pp. 50455052.
[12] Anderson, C. E., Jr., Orphal, D. L., Behner, T., and Templeton, D. W., 2009,
Failure and Penetration Response of Borosilicate Glass During Short-Rod
Impact, International Journal of Impact Engineering, 36, pp. 789798.
[13] Nie, X., Chen, W., Wereszczak, A., and Templeton, D., 2009, Effect of Loading Rate and Surface Conditions on the Flexural Strength of Borosilicate
Glass, J. Am. Ceram. Soc., 92(6), pp. 12871295.
[14] Anderson, C. Jr., Weiss, C., and Chocron, S., 2009, Impact Experiments into
Borosilicate Glass at Three Scale Sizes, Southwest Research Institute, San
Antonio, TX, Technical Report No. 18.12544/018.
[15] Sun, X., Khaleel, M., and Davies, R., 2005, Modeling of Stone-Impact Resistance of Monolithic Glass Ply using Continuum Damage Mechanics, Int. J.
Damage Mech., 14, pp. 165178.
[16] Wereszczak, A. A., Kirkland, T. P., Ragan, M. E., Strong, K. T., Jr., and Lin,
H., 2010, Size Scaling of Tensile Failure Stress in a Float Soda-Lime-Silicate
Glass, International Journal of Applied Glass Science, 1(2), pp. 143150.
[17] Chocron, S., Anderson, C. E., Jr., Nicholls, E., and Dannemann, K. A.,
Characterization of Confined Intact and Damaged Borosilicate Glass, J. Am.
Ceram. Soc. (to be published).
[18] Simha, C. and Gupta, Y., 2004, Time-Dependent Inelastic Deformation of
Shocked Soda-Lime Glass, J. Appl. Phys., 96(4), pp. 18801890.
[19] Sundaram, S., 1993, Pressure-Shear Plate Impact Studies of Alumina Ceramics
and the Influence of an Intergranular Glassy Phase, Ph.D. thesis, Brown University, Providence, RI.
[20] Cagnoux, J., 1982, Shock-Wave Compression of a Borosilicate Glass up to
170 kbar, Shock Compression of Condensed Matter-1982, pp. 392296.
[21] Alexander, C. S., Chhabildas, L. C., Reinhart, W. D., and Templeton, D. W.,
2008, Changes to the Shock Response of Fused Quartz due to Glass Modification, International Journal of Impact Engineering, 35, pp. 13761385.
[22] Handin, J., Heard, H. C., and Magouirk, J. N., 1967, Effects of the Intermediate Principal Stress on the Failure of Limestone, Dolomite, and Glass at Different Temperatures and Strain Rates, J. Geophys. Res., 72(2), pp. 611640.
[23] Chen, W., 2010, Purdue University, private communication.
[24] Glenn, L. A., Moran, B., and Kusubov, A. S., 1990, Modeling Jet Penetration
in Glass, Lawrence Livermore National Laboratory, CA., Technical Report
No. UCRL-JC-103512.
[25] Behner, T., Anderson, C., Jr., Orphal, D., Hohler, V., Moll, M., and Templeton,
D., 2008, Penetration and Failure of Lead and Borosilicate Glass against Rod
Impact, International Journal of Impact Engineering, 35, pp. 447456.
[26] Anderson, C. E., Jr., Behner, Th., Holmquist, T. J., Wickert, M., Hohler, V.,
and Templeton, D. W., 2007, Interface Defeat of Long Rods Impacting Borosilicate Glass, Proceedings of the 23rd International Symposium on Ballistics,
Tarragona, Spain, pp. 10491056.
[27] Johnson, G. R., Holmquist, T. J., and Beissel, S. R., 2003, Response of Aluminum Nitride (Including a Phase Change) to Large Strains, High Strain Rates,
and High Pressures, J. Appl. Phys., 94(3), pp. 16391646.
[28] Frank, A. and Adley, M., 2007, On the Importance of a Three-Invariant Model
for Simulating the Perforation of Concrete Targets, Proceedings from the 78th
Shock and Vibration Symposium, Philadelphia, PA.
[29] Johnson, G. R. and Holmquist, T. J., 1994, An Improved Computational Constitutive Model for Brittle Materials, High Pressure Science and Technology
1993, S. C. Schmidt, J. W. Schaner, G. A. Samara, and M. Ross, eds., AIP,
New York, pp. 981984.
[30] Johnson, G. R., Stryk, R. A., Holmquist, T. J., and Beissel, S. R., 1997,
Numerical Algorithms in a Lagrangian Hydrocode, Wright Laboratory, FL,
Technical Report No. WL-TR-1997-7039.
[31] Johnson, G. R., Beissel, S. R., and Stryk, R. A., 2002, An Improved Generalized Particle Algorithm that Includes Boundaries and Interfaces, Int. J. Numer.
Methods Eng., 53, pp. 875904.
[32] Johnson, G. R., Stryk, R. A., Beissel, S. R., and Holmquist, T. J., 2002,
Conversion of Finite Elements into Meshless Particles for Penetration Computations Involving Ceramic Targets, Shock Compression of Condensed
Matter2001, M. D. Furnish, N. N. Thadhani, and Y. Horie, eds., AIP, New
York.

SEPTEMBER 2011, Vol. 78 / 051003-9

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 11/18/2014 Terms of Use: http://asme.org/terms

Das könnte Ihnen auch gefallen