You are on page 1of 9

Int. J. Appl. Ceram. Technol.

, 7 [3] 400–408 (2010)


DOI:10.1111/j.1744-7402.2009.02355.x

Ceramic Product Development and Commercialization

Failure Waves and their Possible Roles in Determining


Penetration Resistance of Glass
Stephan J. Bless*
Institute for Advanced Technology, The University of Texas at Austin, Austin, Texas 78759

High-velocity impact can produce failure waves in glass. Failure waves limit the speed at which a glass target can become
comminuted. The resulting time dependence of strength is likely to influence resistance to ballistic penetration.

Introduction Failure Fronts Produced by Projectile Impact

In recent years, there has been a large demand for Under high-speed impacts, fracture, and damage
improved transparent armor. The conceptual frame- propagation occur over time scales commensurate with
work for transparent armor design still comes largely the applied loads, leading to new phenomena. For exam-
from investigations of indentation and low-speed im- ple, photographs of impacts at higher speeds into trans-
pacts on glass. The dominant failure sequence under parent brittle materials often show a dark zone that
those conditions is formation of a cone crack that ini- propagates away from the impact site. This zone often
tiates around the periphery of the indentor. Additional stops and evolves into a dense crack array, as shown in
loading produces either classic cone cracks or median Fig. 1. These advancing failure zones are referred to as a
cracks, depending on the amount of ductility. See, for failure wave (FW). The first mention of the concept of a
example, Lawn.1 ‘‘FW’’ apparently occurs in Russian publications dating
from the 1960s, such as Galin et al.,2 Nikolaevskii,3 and
Originally presented at the 46th Sagamore Army Materials Research Conference on Ad- Cherepanov,4 where ‘‘self-propagating failure fronts’’ are
vances and Needs in Multi-Special Transparent Materials Technology in St. Michaels, MD, discussed.
May 9–12, 2005. Supported by the U.S. Army Research Laboratory under contract number Many other observations of impact-induced damage
DAAD17-01-D-0001.
*stephan_bless@iat.utexas.edu can also probably be interpreted as FWs. In Fig. 2, a
r 2009 The American Ceramic Society two-layer glass target is penetrated by a shaped charge jet,
www.ceramics.org/ACT Failure Waves in Determining Penetration Resistance of Glass 401

Fig. 2. Hypervelocity penetration.6 Shaped charge jet enters


layered glass from the right. The opaque zone on the right is a
failure wave (FW) advancing from the impact surface. The arrows
point to a densified region around the cavity. At the time of the
photograph, the jet has entered the second layer, and a new FW has
initiated at the interface.

blunt nose projectile. The conical nose projectile pro-


duces a front of propagating needle cracks, such as have
been recovered from impacts onto glass armor,8 while
the blunt projectile produces a large failed region whose
advance apparently involves nucleation of damage ahead
of the region that is already wholly damaged.
From these and other observations, the properties
of the moving failed zone are usually found to include
the following:
 it is apparently a region where the glass has been
comminuted (broken into very small particles), and
 it moves with a characteristic velocity that often
exceeds the normal maximum crack speed (about
Fig. 1. Impact of sphere on a half-space; comminuted region
1500 m/s for most glasses).
propagates into glass at high speed.5

and the jet has penetrated into the second layer. A failure Properties of FWs Determined by Plate Impact Tests
front is penetrating inward (to the left) from the impact
surface, and also backward from the glass/glass interface. Reports of scientific investigations of FWs began
The speed of this failure front in about 2 km/s. Figure 3 appearing in the early 1990s. The first experiments were
is a side view of a glass block struck by a high-velocity plate impact tests conducted by Gennady Kanel. (Read-
rod. There are two failure regions propagating into the
block, one from the front and one from the rear. Figure 4
shows images from penetration by a tungsten rod at
about 1.4 km/s. An FW that begins with a spherical front
consumes the first of two glass plates. When the projectile
arrives at the interface (Frame 33), a plane FW erupts
from the interface and advances into the second plate.
The above examples were for impacts onto half
spaces, but FWs can also apparently be produced in Fig. 3. Impact of a blunt steel projectile on glass at 700 m/s.
plane stress. Figure 5 shows two images from impacts Projectile strikes top surface. Failure waves are the dark zones that
onto the edge of glass plates—from a conical and a propagate from the impact side and back side.7
402 International Journal of Applied Ceramic Technology—Bless Vol. 7, No. 3, 2010

ers unfamiliar with plate impact test techniques are re-


ferred to one of several recent review publications.10,11)
Kanel communicated his early experiments to the
Impact face
author at a conference in 1990 and later published them
Early failure front in 1991.12 Figures 6a and b are from Kanel’s original
figures. Figure 6a is an (x, t) diagram for impact of a
Interface between copper flyer plate onto a glass target that occupies the
two touching space x40. Elastic and deformational shocks propagate
25 mm glass plates into the glass sample. Kanel was looking for a reflection
back from the front of the flyer plate. Instead, he ob-
served an unexpected arrival that apparently had re-
flected from an interface within the target, shown as the
inner dotted line in Fig. 6a. Experiments with different
thickness samples established that the mysterious
interface was moving (Fig. 6b). Bless et al.13and his
colleagues at the University of Dayton did similar
experiments and obtained consistent results.
Fig. 4. Impact of a tungsten rod onto two glass plates at about
1.4 km/s. Projectile moves right to left. Frame 1 is just after impact. Subsequently, the team at Dayton conducted spall
Frame 33 is just after the projectile enters the second plate. stress measurements on glass. Spall stress is also mea-

Fig. 5. Two images from impacts onto the edge of glass plates—from a conical and a blunt nose projectile.9
www.ceramics.org/ACT Failure Waves in Determining Penetration Resistance of Glass 403

Fig. 6. (a) Diagram of original failure wave (FW) experiment by


Kanel et al.12 (b) Kanel et al. results.

sured in the plate-impact configuration, and corre-


sponds to the bulk tensile strength of a material (e.g.,
tensile failure unaffected by surface flaws). They found
that the spall stress of soda lime glass in front of the FW
was extremely high—above 6 GPa. However, the spall
strength behind the FWs was nearly zero. This demon-
strated that the failure front separates glass that is intact
from glass that has been comminuted. Soon afterward, Fig. 8. Photographs reveal that fronts often are not smooth.16 The
the Dayton team also measured the shear strength in failure wave moves from left to right. Time after impact is marked
on the photographs. A wave shock (S) is visible in the first and
front of and behind the FW. Their results are shown in
second frames.
Fig. 7; ahead of the failure front, the shear stress was
related to the principal stress by the elastic relationship,
but behind the failure front the shear stress relaxed to a
value a little more than 20 kbar (2 GPa). Similar mea-
surements have now been made on many other types of trated in Figs. 1, 4, and 5 (left), the wave fronts are not
glass.14 smooth. They apparently advance by developing
Several investigators have also photographed plane ‘‘fingers’’ that progress just ahead of the main front, as
FW fronts. As was the case for projectile impacts illus- illustrated in Fig. 8 from a plate impact test. The
instability of propagating fractures such as these is
predicted by a recent theoretical analysis.15
Interfaces have a significant affect on FW propaga-
tion. This was observed in projectile impact studies, and
also in plate impacts. It appears that FWs are generated
from interfaces, as in Fig. 9, taken from Kanel et al.17 In
these experiments, as the shock wave crossed each in-
terface, FWs were generated that propagated both in the
forward direction, following the shock, and in the back-
ward direction. Failure waves are driven by stress
fields—when the stress is removed by release waves ar-
riving from free surfaces or by exhaustion of the pro-
jectile,18 they stop.
Fig. 7. Stress deviator decreased—showing decrease in shear Observations of FWs as revealed by plate impact
strength behind FW. HEL, Hugoniot elastic limit. tests over the past decade may be summarized as follows:
404 International Journal of Applied Ceramic Technology—Bless Vol. 7, No. 3, 2010

Other Examples of FWs

Prince Rupert’s drops are one FW example that has


often been used in classroom demonstrations. They are
formed when liquid glass drops are quenched by drop-
ping them into water. As a consequence, the outside is in
compression, while the interior is under tension. When
the tail is broken, these drops rapidly explode. The driv-
ing mechanism is apparently twofold: there is a great
deal of stored elastic energy in the drop that is available
to drive fragments, and glass is dilatant. Figure 10 illus-
trates an exploding Prince Rupert’s drop.
Fig. 9. Diagram for copper striking glass layers. E represents Another geometry in which FWs are observed is bars
shocks. Double lines are release waves from closing the small gap or prisms. Figure 11 shows an FW in a glass bar that has
between layers. Failure wave (FWs) are generated at interfaces.17 been struck on one end face by a high-speed projectile.
The FW starts at the projectile interface and rapidly con-
sumes the entire visible bar. The speed of FWs in bars
 There is a threshold stress—about half the elastic seems to approach O2 times the shear wave speed.
limit—for formation of FWs. (The elastic limit It is plausible that FWs can occur in tension, al-
is not shear failure in glass; it is apparently dens- though this has not been shown conclusively. For ex-
ification.) ample, Fig. 12 shows the FW propagating through the
 Mechanical properties change little in the longi- tensile region of a Prince Rupert’s drop. Figure 13
tudinal direction across the FW. shows a glass bar in which the distil end explodes in a
 Transverse stress increases and shear stress de- tensile failure. It is interesting that the fractured material
creases at the FW front—approaching that of a seems to differ in these two experiments. The Prince
granular material. Rupert’s drop produces platelets, whereas the bar pro-
 There is an increase in mean stress (e.g., dilatancy or duces faceted, nontransparent nuggets.
bulking). (This comes about because glass normally
expands when it fractures, because open cracks take
up volume. In these experiments, the glass is frac- Theory
tured at constant volume, so it self pressurizes.)
 FW have been observed in many different There have been several attempts to model FWs,
glasses, including fused silica. none entirely satisfactory. Most models are phenome-
 Material behind the FW has little or no tensile nological, meaning that the behavior is assumed and
strength. then described mathematically. The decrease in the
 FW speeds can exceed the shear wave speed. stress deviator can be modeled as a drop in shear mod-
They greatly exceed crack velocities. Speed often ulus11,21 or a decrease in strength.22
decreases with decreasing driving stress. Feng23 has modeled the FW as crack diffusion with
 FW fronts are often not smooth. some success. Naimark24 describes FWs as a possible
 FWs can start at interfaces. They also can stop at collective behavior mode of defects under very short
interfaces. duration loading. Simha and Gupta25 consider FWs to
 FWs are quenched by release waves. be an example of delayed fracture that is instigated by
the arrival of the primary stress wave.15
It should also be pointed out that FWs are outside The most compelling numerical modeling of a FW
the framework of all current models for penetration of concerned the bar impact geometry, and was accom-
brittle materials. Until the question of the role played by plished by Repetto et al.26 In a very high-resolution
FWs in ballistic transparencies is resolved, understand- calculation of individual cracks, they reproduced the
ing derived from theoretical or numerical treatments FW evolution in glass bars. Cracks initiate according to
must, therefore, be viewed cautiously. a tensile failure criterion. The time dependence of the
www.ceramics.org/ACT Failure Waves in Determining Penetration Resistance of Glass 405

Fig. 10. An exploding Prince Rupert’s drop.5

FW was controlled by a cohesive law in which strength crystalline ceramics, chiefly alumina, SiC, or boron car-
decreased as a function of crack opening displacement. bide. Conventional ceramic armor is composed of tile
mosaics. The armor is normally designed so that impact
damage remains confined to the impacted tile, and in
Implications for Armor this way multiple-hit performance is achieved. Trans-
parent armor differs from these conventional ceramic
Transparent armor is an example of ceramic armor. armors; it is normally composed of several layers, which
Opaque ceramic armor is normally comprised of poly- is the means by which multihit performance is achieved.
Direct observation of FWs in glass laminate armor is
reported in.8
Against this background, one can speculate on the
possible role of FWs for transparent armor. For thin
glass layers, one would expect an enhanced vulnerability
to blunt projectiles, which produce larger FWs (e.g.,
Fig. 3). In addition, there should be a critical velocity
below which the armor is more effective because there is
no FW. When an FW occurs, a significant volume of
material is transformed into a powder around the im-
pact site. This material offers very little penetration re-
sistance. In a multilayer glass target, it is probable that
much of the penetration resistance comes from interior
layers that have not been consumed by a failure wave.
When there are no interlayers to arrest the motion of
Fig. 11. Failure wave (FW) in a glass bar, with impact in Frame FWs, most of the penetration takes place in material
2. In Frame 4, the wave is halfway across the field of view. In that has been comminuted by the FW (e.g., Anderson
Frame 6, the bar has been consumed.19 et al.27) The decrease in penetration resistance as the
406 International Journal of Applied Ceramic Technology—Bless Vol. 7, No. 3, 2010

Fig. 12. Failure wave in Prince Rupert’s drop and recovered particle (inset).5

target is consumed by an FW is also shown in Fig. 14. The case for sharp projectiles is frankly much more
Most of the penetration takes place in the low-strength speculative. The region of high compressive stress asso-
material. ciated with a sharp impact is very limited. Failure waves
It also follows that there is a penetration velocity from the back side of glass plates (provided they are not
above which the projectiles would move supersonic rel- too thick) would be expected. Thus, glass may offer an
ative to the FW. Above this velocity, the glass is more initially very high resistance to sharp bullets, but for
difficult to penetrate. This has been demonstrated in a relatively thin tiles, the resistance of the glass may only
recent U.S.–Russian collaboration.29 The critical veloc- last until the plate is consumed by the tensile failure
ity is well above those normally considered as transpar- wave. These processes may explain the fact that the tips
ency threats. of hard bullets are often broken by glass, but (unlike the

Fig. 13. Float glass bar struck by projectile on left and recovered Fig. 14. Effective strength of glass as a function of time after
particle from distil end (inset).20 impact.28
www.ceramics.org/ACT Failure Waves in Determining Penetration Resistance of Glass 407

case of ceramic armor), the body of the bullet suffers Research Laboratory or the U.S. Government unless so
relatively little damage and is often turned sideways, as designated by other authorized documents. Citation of
happens in sand penetration. manufacturer’s or trade names does not constitute an
One hopes that better understanding of FWs will official endorsement or approval of the use thereof. The
enable the design of better transparent armor. How this U.S. Government is authorized to reproduce and dis-
could come about is of course speculation at this time; tribute reprints for government purposes notwithstand-
however, perhaps the following concepts might be useful: ing any copyright notation hereon.
1. Suppression of FWs: Cavity expansion models
predict that superposition of compressive stress can sup-
press FWs.30 This might be achieved by surface treat-
References
ments, which now can be as high as 1 GPa.31
2. Ultra-pure materials: These should have 1. B. Lawn, Fracture of Brittle Solids, 2nd Edition, Cambridge University Press,
higher FW thresholds and should be examined for this Cambridge, U.K., 1998.
effect. 2. L. A. Galin, and G. P. Cherepanov, ‘‘Self-Sustaining Failure of a Stressed
Brittle Body,’’ Sov Physics – Doklady, 11 267–269 (1966).
3. Use of interlayers to suppress FWs: It is clear that 3. V. N. Nikolaevskii, ‘‘Limit Velocity of Fracture Front and Dynamic Strength
FWs can stop or be generated at interfaces. Suppression of Brittle Solids,’’ Int J. Eng. Sci., 19 41–56 (1981).
4. G. P. Cherepanov, Mechanics of Brittle Failure. A. P. Peabody, transl. Mc-
might greatly benefit transparent armor. The criteria Graw-Hill, New York, 536–540, 1979.
have not been well investigated. 5. S. Chandrasekar and M. M. Chaudri, ‘‘The Explosive Disintegration of
Prince Rupert’s Drops,’’ Phil. Mag. B, 70 [6] 1195–1218 (1994).
4. New materials: Transparent polycrystals, single 6. G. E. Hauver, P. H. Netherwood, R. F. Benck, and A. Melani, ‘‘Penetration
crystals, and glass ceramics, will soon be available. These of Shaped-Charge Jets into Glass and Crystalline Quartz,’’ U.S. Army Bal-
listic Research Laboratory Report BRL-TR-3273f, 1991.
materials should be investigated to determine if they 7. A. S. Vlasov, E. I. Zilberbrand, A. A. Kozhushko, A. I. Kozachuk, A. B. Si-
support FWs. nani, and M. I. Stepanov ‘‘Ballistic Behavior of Strengthened Silicate Glass,’’
20th International Symposium on Ballistics, Orlando, FL, September 23–27,
2002.
8. S. J. Bless, T. Chen, and R. Russell ‘‘Impact on Glass Laminates,’’ 23rd In-
Summary ternational Symposium on Ballistics (ISB), April 16–20, 2007, Tarragona,
Spain.
9. U. Hornemann, H. Rothenhäusler, J. F. Senf, S. Kalthoff, and S. Winkler,
Failure and penetration of glass by high-speed pro- ‘‘Experimental Investigation of Wave and Fracture Propagation in Glass Slabs
jectiles differs from static indentation in several impor- Loaded by Steel Cylinders at High Impact Velocity,’’ Inst. Pys. Conf. Ser., 70
tant ways. Damage can occur not by single cracks, but 291–298 (1984).
10. G. I. Kanel, S. V. Razorenov, and V. E. Fortov, Shock-Wave Phenomena and
by advancing networks of cracks that reduce the glass to the Properties of Condensed Matter, Springer, New York, 2004.
rubble. The failure zones possess finite propagation 11. M. A. Meyers, Dynamic Behavior of Materials, Wiley, New York, 1994.
12. G. I. Kanel, S. V. Rasorenov, and V. E. Fortov, ‘‘The Failure Waves and
speeds that can render the time scale for glass fracture Spallations in Homogeneous Brittle Materials,’’ Shock Compression of Con-
commensurate with the impact time scale, and this may densed Matter—1991, Proceedings of the American Physical Society Topical
Conference on Shock Compression of Condensed Matter, June 17–20, 1991. eds.,
have a profound effect on the ability of the glass to stop S. C. Schmidt, et al., Elsevier Science Publishing, Williamsburg, VA,
the projectile. Understanding how failure waves are gen- 451–454, 1992.
13. S. J. Bless, N. S. Brar, G. Kanel, and Z. Rosenberg, ‘‘Failure Waves in Glass,’’
erated and propagated will probably significantly ad- J. Am. Ceram. Soc., 75 1002–1004 (1992).
vance the scientific understanding required for efficient 14. S. Bless and N. S. Brar, ‘‘Failure Waves and Their Effects on Penetration
Mechanics in Glass and Ceramics,’’ Chapter 3 in Shock Wave Science
design of transparent armor. and Technology Reference Library, ed., Y. Horie, Springer-Verlag, Berlin,
2007.
15. M. A. Grinfeld, S. E. Schoenfeld, and T. W. Wright, ‘‘Morphological Insta-
Acknowledgment bility of Failure Fronts,’’ Appl. Phys. Lett., 88 104102 (2006).
16. N. K. Bourne and Z. Rosenberg, ‘‘The Dynamic Response of Soda-
Lime Glass,’’ Shock Compression of Condensed Matter—1995, 370. eds., S.
The research reported in this document was per- C. Schmidt and W. C. Tao. AIP Conference Proceedings, Woodbury, NY,
formed in connection with Contract number DAAD17- 567–572, 1996.
17. G. I. Kanel, A. A. Bogatch, S. V. Razorenov, and Z. Chen, ‘‘Transformation
01-D-0001 with the U.S. Army Research Laboratory of Shock Compression Pulses in Glass Due to the Failure Wave Phenomena,’’
and Award number N00014-06-1-0475 with the Office J. Appl. Phys., 92 5045–5052 (2002).
18. C. Anderson and D. Orphal, ‘‘Response of Glass to Short Rod Impact. Shock
of Naval Research. The views and conclusions contained Compression of Condensed Matter–2007,’’ Am. Phys. Soc., (2008).
in this document are those of the authors and should 19. S. J. Bless, N. S. Brar, G. Kanel, and Z. Rosenberg, ‘‘Failure Waves in Glass,’’
J. Am. Ceram. Soc., 75 1002–1004 (1992).
not be interpreted as presenting the official policies or 20. P. Zeinert, S. Bless, and R. Nichols, ‘‘Comminuted Particles Originating
position, either expressed or implied, of the U.S. Army from Catastrophic Failure of a Glass Bar,’’ Shock Compression of Condensed
408 International Journal of Applied Ceramic Technology—Bless Vol. 7, No. 3, 2010

Matter-2005, eds., M. D. Furnish, M. Elert, T. P. Russell, and C. T. White, 27. C. E. Anderson, I. S. Chocron, and J. D. Walker ‘‘Analysis of Time-Resolved
AIP Conference Proceedings 845, 900–902. AIP Melville, NY, 2007. Penetration of Long Rods into Glass Targets,’’ Ceramic Engineering Science
21. Y. Partom, ‘‘Modeling Failure Waves in Glass,’’ Int. J. Impact Eng., 21 791– Proceedings, V. 26, n7, Advances in Ceramic Armor, A collection of Papers
799 (1998). Presented at the 29th International Conference on Advanced Ceramics and
22. H. D. Espinosa, Y. Xu, and N. S. Brar, ‘‘Micromechanics of Failure Waves in Composites, 27–34, 2005.
Glass II: Modeling,’’ J. Am. Ceram. Soc., 80 2074–2085 (1997). 28. A. A. Kozhushko and G. S. Pugachev, unpublished data.
23. R. Feng, ‘‘Formation and Propagation of Failure in Shocked Glasses,’’ J. Appl. 29. E. L. Zilberbrand, A. S. Vlasov, J. U. Cazamias, S. J. Bless, and A. A.
Phys., 87 1693–1700 (2000). Kozhushko, ‘‘Failure Wave Effects in Hypervelocity Penetration,’’ Int. J. Im-
24. O. B. Naimark, ‘‘Collective Properties of Defect Ensembles and Some Non- pact Eng., 23 95–1001 (1999).
linear Problems of Plasticity and Fracture,’’ Phys. Mesomechan., 6 39–63 30. S. Satapathy, S. Bless, and S. Ivanov ‘‘The Effects of Failure Waves on Pen-
(2003). etration Resistance of Glass,’’ 18th International Symposium Ballistics Sympo-
25. C. H. Simha and M. Gupta, ‘‘Time-Dependent Inelastic Deformation of sium, San Antonio, TX, November 1999.
Shocked Soda–Lime Glass,’’ J. App. Phys., 96 1880–1890 (2004). 31. A. K. Varhneya, W. C. LaCourse, and I. Spinelli ‘‘Faster and Deeper Chem-
26. E. A. Repetto, R. Radovitzky, and M. Ortiz, ‘‘Finite Element Simulation of ical Strengthening of Glass for Security Applications,’’ 46th Sagamore Army
Dynamic Fracture and Fragmentation of Glass Rods,’’ Comput. Methods Appl. Materials Research Conference on Advances and Needs in Multi-Special Trans-
Mech. Eng. 183 3–14 (2000). parent Materials Technology, St. Michaels, MA, May 9–12, 2005.

You might also like