Dynamic Fracture and Fragmentation of Boron Carbide Advanced Ceramics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 100

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/338420263

DYNAMIC FRACTURE AND FRAGMENTATION OF BORON CARBIDE ADVANCED


CERAMICS

Chapter · January 2020

CITATION READS

1 986

5 authors, including:

Haoyang Li Calvin Lo
University of Alberta University of Alberta
20 PUBLICATIONS   38 CITATIONS    8 PUBLICATIONS   19 CITATIONS   

SEE PROFILE SEE PROFILE

James D. Hogan
University of Alberta
101 PUBLICATIONS   685 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Study the mechanical behaviors and failure mechanisms of the TiAl/Ti3Al/Al2O3 cermet under high strain rates View project

All content following this page was uploaded by James D. Hogan on 07 January 2020.

The user has requested enhancement of the downloaded file.


Chapter

DYNAMIC FRACTURE AND FRAGMENTATION


OF BORON CARBIDE ADVANCED CERAMICS

Haoyang Li1, Calvin Lo1, Geneviève Toussaint2, Tomoko


Sano3, James D. Hogan1*

1. Department of Mechanical Engineering, the University of Alberta,


Edmonton, AB T6G 2R3, Canada
2. Defence Research and Development Canada, Valcartier Research
Center, 2459 de la Bravoure Road, Québec, QC, G3J 1X5, Canada
3. Weapons and Materials Research Directorate, Army Research
Laboratory, Aberdeen Proving Ground, MD 21005

ABSTRACT
This chapter reviews the different mechanical testing, characterization
approaches, and computational models used to study fracture and
fragmentation of boron carbide advanced ceramics, highlighting
microstructural dependencies of these behaviors. Boron carbide, as an
advanced ceramics with low density, high hardness and stiffness, and strong
abrasion resistance, has attracted significant engineering research attention in
recent decades for structural applications, such as light-weight armors, high-
temperature coatings, and in blasting nozzles. In these applications, impacts
will eventually lead to fracture and then fragmentation of the material. As a
consequence, studying the fracture and fragmentation of boron carbide
becomes crucial in understanding the material performance and determining
its functionality under dynamic extreme environments (i.e., high stress, high
loading rate, and high temperature). Throughout the years, mechanical testing
(e.g., split-Hopkinson pressure bar) coupled with real-time visualization
techniques (e.g., high-speed imaging) at different strain rates and stress states
have been employed to study the in-situ fracture and fragmentation of boron

* Corresponding
Author address
Email: jdhogan@ualberta.ca
2 Li et al.

carbide. More recently, advancements in image processing, digital image


correlation, virtual fields method, and micro-computerized tomography have
been integrated into different testing setups in order to acquire more
comprehensive data sets of fracture and fragmentation. Postmortem analysis
techniques, such as advanced microscopy (e.g., scanning electron microscopy
and transmission electron microscopy), and automated morphological
investigation of particles after testing have also been used to study the fracture
and fragmentation mechanisms of boron carbide. In conjunction with
advancements in mechanical testing and characterization technologies,
computational models have also been developed to predict the mechanical
responses and failure of boron carbide and other advanced ceramics. These
approaches include phenomenological models (e.g., Johnson-Holmquist-2
model), micro-mechanical models (e.g., wing crack model), and fragmentation
models (e.g., process-driven energy-based model). Altogether, this chapter
will review recent progress in understanding fracture and fragmentation of
boron carbide through experimentation, characterization, and modeling. The
chapter concludes with promising future directions in these approaches.

Keywords: boron carbide, fracture, fragmentation, mechanical testing, impact


testing, modeling, mechanism, characterization, micro-mechanical,
phenomenological, brittle, ceramic

1. INTRODUCTION
The low density (~ 2.5 g/cm3, manufacturing method dependent) (Roy,
Subramanian, and Suri 2006), high hardness (~ 25 to 30 GPa, grain size dependent)
(Vargas-Gonzalez, Speyer, and Campbell 2010), and high strength (~ 3 to 5 GPa,
manufacturing and grain size dependent) (Swab et al. 2017) of boron carbide make
it a light-weight structural material that is ideal in ceramics-based armour systems.
As a result, much research has been focused on the dynamic behavior of boron
carbide (e.g., Gooch et al. 2000; Vogler, Reinhart, and Chhabildas 2004; Ghosh et
al. 2007; Clayton 2013). More so than in other structural applications of materials,
the fracture and fragmentation behavior of advanced ceramics is central to its
ballistic performance. Impact events are characterized by the deposition of large
amounts of energy in an extremely short time. The fracture and fragmentation in
brittle materials, such as boron carbide, are effective ways to dissipate the impact
energy by generating new fracture surfaces. As noted by (Wilkins 1968), “If an
armor doesn’t fail for a given ballistic threat, it could be made lighter.” Therefore,
material failure will always be an important consideration for armor that includes
boron carbide as a constituent material. In addition, fragmentation, which is
considered as the last stage of brittle material failure, has been shown to be critical
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 3

to the performance of ceramic body armor (e.g., Moynihan, LaSalvia, and Burkins
2002; Krell and Strassburger 2014), as fragmentation behavior is linked to
projectile erosion. For these reasons, the development of next-generation boron
carbide armor requires an improved understanding of the fracture and
fragmentation behavior of boron carbide under complex stress states and dynamic
loading conditions.
Fracture and fragmentation mechanisms are generally stress-state and strain-
rate dependent. Consequently, a range of laboratory-based experiments has been
developed to access different stress-states and strain-rates under controlled and
well-defined conditions in order to study important fracture and fragmentation
behaviors. For example, the split-Hopkinson pressure bar systems, which was first
developed by (Kolsky 1963) and later modified to study various stress states, have
been used to access strain rates between 102 s-1 and 104 s-1 on boron carbide (e.g.,
Paliwal and Ramesh 2007; Farbaniec et al. 2016; Sano et al. 2018). For even higher
strain rates (from 104 s-1 to 107 s-1), ballistic impact, plate impact, laser spall, and
laser shock experiments have been employed to probe the fracture and
fragmentation responses of boron carbide at these extreme conditions (e.g.,
Holmquist and Johnson 2006; LaSalvia et al. 2009; Taylor 2015; Cui, Ma, and Li
2017). More recently, state-of-the-art real-time visualization techniques (e.g., ultra-
high-speed imaging) have been coupled with these experimental setups to explore
the in-situ fracture and fragmentation behaviors of boron carbide during dynamic
loading (e.g., Hogan et al. 2015; Crouch, Appleby-Thomas, and Hazell 2015). In
addition, advanced characterization techniques (e.g., high-resolution transmission
electron microscope) have been utilized post-mortem to investigate the
microstructure of recovered fragments and explore the connections between
microstructure and failure mechanisms (e.g., Domnich et al. 2011).
Predictive models incorporate the insights and data derived from the controlled
conditions of laboratory experiments to bridge the gap between experimental
studies and the complex loading conditions of large-scale ballistic impacts. Fracture
and damage models, including phenomenological (e.g., Johnson-Holmquist-2
model (G. R. Johnson and Holmquist 2008)) and micro-mechanical models (e.g.,
the wing crack model (Nemat-Nasser and Horii 1982)) are often used to predict the
mechanical response (i.e., stress-strain behavior, strength degradation) and failure
(i.e., crack network, plastic deformation) of advanced ceramics (e.g., Rajendran and
Kroupa 1989; Simha, Bless, and Bedford 2001) and boron carbide (e.g., Clayton
2013; Clayton 2014;) under specified stress states and strain rates.
Phenomenological models describe the empirical relationships between loading
conditions and the changes in mechanical properties. These models rely on data
obtained from experiments to calculate model parameters, and these models can be
4 Li et al.

run efficiently for conditions bounded within the data set used for calibration.
Alternatively, micro-mechanical models are developed to incorporate
microstructural features (e.g., grain size, secondary phase, and defect size and
shape) and important failure mechanisms in order to investigate brittle material
responses (Ju, Lee, and Member 1991; Curran et al. 1993; Pensée, Kondo, and
Dormieux 2002; Zhu and Tang 2004). In practice, both modeling approaches have
been implemented into large-scale computational simulations, where material
fracturing during an impact event can be visualized (Tonge and Ramesh 2016a,
2016b). Beyond large-scale modeling used to simulate the entire impact event,
fragmentation modeling has been centered on predicting average fragment sizes
and the associated distribution of sizes. The bulk of these works have focused on
the fragmentation generated from tensile states (e.g., Mott 1945; Grady 1982;
Grady and Kipp 1985a; Glenn and Chudnovskly 1986; Zhou, Molinari, and Ramesh
2006a, 2006b; Grady and Mott 2006; Levy and Molinari 2010). These models
evolved from energy-based analytical models, which considered idealized
geometry and used statistical data (e.g., Grady and Mott 2006) that integrated the
rate-dependent fragment size and localized cohesive zone for energy dissipation
(e.g., Zhou, Molinari, and Ramesh 2005). These models are important because they
can be integrated into impact failure models to better predict fragmentation
consequences of impact.
This chapter is organized as follows: First, various mechanical testing methods
which are used to explore the fracture and fragmentation of boron carbide at
different stress states and strain rates will be presented. The focus will be on the
dynamic loading regime, which is commonly accessed in the intended applications
of the material (e.g., impact). Key experimental results and micrographs identifying
fracture and fragmentation mechanisms will be shown together with each
mechanical testing method. Through this first part, the readers should become
familiarized with the state-of-the-art testing methods which have been used to study
boron carbide and other brittle materials. This should give insights for the
researchers who work in the relevant fields. Next, computational modeling
frameworks on the fracture and fragmentation of advanced ceramics and boron
carbide will be presented along with critical simulation outcomes. This work is
necessary because models can be used to better understand experimental outcomes
and, more importantly, guide further design of boron carbide materials and
protective structures incorporating the materials. Finally, we conclude with a
discussion of future directions for studying the failure of boron carbide, followed
by remarks summarizing the chapter.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 5

2. STUDYING THE FRACTURE AND FRAGMENTATION OF


BORON CARBIDE USING MECHANICAL AND IMPACT TESTING
In this section, the experimental apparatuses used to investigate the fracture and
fragmentation of boron carbide under various stress states and strain rates will be
described. Many of these setups will be coupled with schematics and key
experimental results taken from the literature. Where applicable, representative
fracture and fragmentation mechanisms in boron carbide explored through
microscopy or spectroscopy will be discussed in conjunction with their
corresponding testing methods.

2.1 The split-Hopkinson pressure bar System

The split-Hopkinson pressure bar (SHPB), which was first developed by


Kolsky (Kolsky 1963), is widely used in studying the dynamic failure of different
groups of materials, including soft material (e.g., foams (Bhagavathula et al. 2018)),
metals (Chen et al. 2003), and ceramics (Chen and Ravichandran 1997). There has
been limited SPHB testing on boron carbide (e.g., Paliwal and Ramesh 2007;
Farbaniec, Hogan, and Ramesh 2015; Hogan et al. 2015; DeVries et al. 2016;
Farbaniec et al. 2017). Generally, the accessible strain rates for an SHPB are
between 102 s-1 and 104 s-1. In order to study brittle materials, the conventional
SHPB system has to be modified (e.g., pulse shapers, platens, momentum traps)
(Chen and Song 2010) so that the material responses can be well captured. The
theory of SHPB system is omitted here, and the details can be referred to in (Chen
and Song 2010).

2.1.1 The Dynamic Uniaxial and Confined Compression Testing using SHPB

Shown in Figure 1 is an example of the modified version of the SHPB system


used to study the dynamic uniaxial compressive behavior of advanced ceramics.
The conventional SHPB system was developed to characterize metals, so
modifications were made to accommodate the high hardness and brittle nature of
ceramics. Some critical components include: 1. The pulse shaper, for which the
size, shape, and thickness need to be carefully chosen to create a ramp-up profile
and a reasonable rise time; 2. The tungsten carbide (WC) platens need to be
impedance matched with the bars to minimize wave reflections at interfaces; 3.
Specimen size and shape need to be designed to mitigate the specimen size effect
(Tekalur and Sen 2011). Unlike metals and soft materials, only the stress in the
6 Li et al.

ceramic specimen can be determined through the transmitted signal captured using
the strain gauge. The expression of the nominal stress in the specimen is:

𝐴0
𝜎(𝑡) = 𝐸0 𝜀 (𝑡) (1)
𝐴𝑠 𝑡

where 𝐴𝑠 (m2) is the cross-sectional area of the specimen, 𝐴𝑜 (m2) is the cross-
sectional area of the bar, and 𝜀𝑡 is the time-resolved transmitted strain. Note that
Equation (1) assumes that the specimen is deformed homogeneously with
equilibrated stress and constant strain rate (Chen and Song 2010), and the bar
deforms elastically during loading. Details on the setup in Figure 1 can be referred
to in (Li, Motamedi, and Hogan 2019).

Figure 1: Schematic of the modified Split-Hopkinson Pressure Bar (SHPB) experimental


setup. This image is taken from (Li, Motamedi, and Hogan 2019).

Different modifications to the SHPB system have been made to achieve multi-
axial loading: 1. Using shrink-fit metal sleeves on cylindrical specimens (Chen and
Ravichandran 2000); 2. Designing a biaxial or triaxial SHPB system (Chen et al.
2018); 3. Sandwiching the specimen between two high-strength steel T-blocks
(Farbaniec et al. 2017). The set-up of (Farbaniec et al. 2017) is shown Figure 2.
Four screws were tightened, each with controlled torques, to generate a
confinement stress of about 500 MPa along the X2 direction. To reduce stress
concentration between the specimen and T-blocks, AISI 4140 low-alloy steel
‘cushions’ of 0.5 mm thickness were used. This setup provided a biaxial stress state
on the boron carbide specimen during compression. The differences in the fracture
and fragmentation behaviors, when compared with the uniaxial case, can then be
assessed, as was done in (Farbaniec et al. 2017).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 7

Figure 2: Confinement setup used in (Farbaniec et al. 2017) for applying a biaxial stress
state on the boron carbide specimen during SHPB testing.

In systems like these, ultra-high-speed imaging is often coupled with an SHPB


setup to obtain real-time visualization of the fracture and fragmentation behaviors
of boron carbide. This is used to gain insights into cracking, such as initiation sites,
direction of propagation, crack interaction, and crack speeds (e.g., Hogan et al.
2016). The onset and evolution of fragmentation can also be obtained through high-
speed imaging, but the quantitative fragmentation analysis needs to be determined
post-mortem because the deformation causes the material to go out of camera focus
(Hogan et al. 2016; Hogan et al. 2017). This information is then correlated with the
stress history curve for further examining the fracture-strength relations.
Shown in Figure 3 is an example of a uniaxial compression test on a hot-pressed
boron carbide using the SHPB testing described in (Hogan et al. 2015). The stress-
time history curve (on the left) is demonstrated together with the time-resolved
high-speed video images (on the right). The images are selected to be 2 µs apart
from each other and the corresponding times are indicated on the stress-time curve.
The stress rate is 200 MPa/µs in this case, which corresponds to approximately 465
s-1 in strain rate. The peak stress (referred to as the “dynamic compressive strength”)
is 3.90 GPa in this case. The fracture behavior of boron carbide is discussed with
high-speed images. Axial cracking is observed prior to the peak stress (t1) and no
further damage is observed at the peak stress (t2). Crack growth and interaction
mechanisms are identified after the peak stress, including edge failure, transverse
cracking, later-time axial cracking, and crack coalescence. The coalescence of axial
and transverse cracks results in the generation of fragments that are between 830
and 1600 µm in size at time t6. Measurements of velocity of the axial cracks range
8 Li et al.

from approximately 1800 to 2400 m/s (ten total measurements across multiple tests)
with an average of 2000 ± 300 m/s.

Figure 3: Stress-time history of dynamic uniaxial compression of boron carbide with time-
resolved high-speed video images showing mesoscale failure mechanisms (Hogan et al. 2015).
The loading direction is horizontal. The dashed line is a linear fit of 10% and 90% of the peak
stress.

Figure 4 shows another example comparing the uniaxial and confined fracture
and fragmentation behaviors of the same boron carbide using the SHPB setup that
was described in Figure 2. The stress-time history profile (left) and the high-speed
video images (right) corresponding to six different times on the stress-time curve
are used to study the time-dependent fracture and fragmentation of boron carbide
related to confinement. Clear differences in stress build-up and failure modes can
be observed between uniaxial and confined loading conditions. It is speculated that
the confinement affects the reflected stress waves from free specimen surfaces, and
consequently, the bulking rate of the damaged material is different from the uniaxial
condition. In other words, the applied confinement limits the crack opening process
and keeps the material as “solid-like,” which provides a favorable condition for
stress wave interaction. This eventually leads to a higher peak stress with a faster
stress build-up. Regarding fracture behavior, the uniaxial compression (Figure 4-b)
produces column-like fragments through mostly axial cracking followed by the
coalescence with the transverse cracks, also observed in Figure 3 by (Hogan et al.
2015). As for the fracture behaviors under confined condition (see Figure 4-c), no
axial splitting failure mode is observed. What has changed is the crack development
in two directions (toward the main compression axis and confinement axis) is
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 9

almost simultaneous, inducing a mixed-mode of cracking. A mixed-mode of


fragment shapes and sizes are observed after material failure.

Figure 4: (a) Stress-time curves in uniaxial and confined dynamic tests; (b) Photographs
from the high-speed camera showing the dynamic failure process in uniaxial compression; (c)
Photographs from the high-speed camera of the dynamic failure process in confined dynamic
compression. Blue arrows indicate the direction of propagation of the compression waves. The
confinement axis is indicated by green arrows. White arrows indicate the formation and
extension of surface cracks. Specimens in this study measure 3.5 mm x 4mm x 5.3 mm. These
images are taken from (Farbaniec et al. 2017).

Post-mortem analysis, including microscopy investigations on the fracture


mechanisms and statistical examination of the fragments, are often performed to
probe the exact failure modes of boron carbide (Hogan et al. 2016). For example,
scanning electron microscopy (SEM) and transmission electron microscopy (TEM)
micrographs shown in (Farbaniec et al. 2017) identify wing crack formation and
transgranular fracture being the major fracture mechanisms under dynamic uniaxial
compression (shown in Figure 5-a and b), and the initiation sites are often the
carbon inclusions introduced through manufacturing. TEM micrographs of
fragments obtained from confined experiments (see Figure 5-c and d) show strong
grain boundaries with no cavitation damage and no stress-induced dislocations or
deformation twins. This also suggests transgranular fracture as the dominant
fracture mechanism under confined condition. Figure 5-b shows cleavage steps,
10 Li et al.

which likely appear as a consequence of the crack propagation through neighboring


grains with highly twisted boundaries.

Figure 5: SEM micrographs showing: (a) Wing crack formation from the carbon
inclusion; (b) Crack interaction and coalescence leading to transgranular fracture and
structural failure of the sample. ‘C’ stands for carbon inclusion. Blue arrows indicate the
approximate direction of the compressive load; (c) Cleavage steps as a consequence of the
crack propagation through neighboring grains with highly twisted boundaries; (d) Internal
grain structure of the fragment that was polished to electron transparency. These images are
taken from (Farbaniec et al. 2017).

Quantitative fragmentation analysis was performed in (Hogan et al. 2015) to


relate the microstructural features to the fragmentation behavior of a hot-pressed
boron carbide after similar uniaxial and biaxial dynamic compression tests. Shown
in Figure 6-a is the as-received boron carbide microstructure showing defects and
inclusions, and Figure 6-b demonstrates a MATLAB process for isolating the
graphite inclusions and computing the defect spacing. Figure 6-c and d show a
correlation between the fragment size and circularity for both dynamic uniaxial and
confined compressions. Two fragmentation regimes were identified based on size,
where one is considered microstructure-dependent (i.e., results in smaller fragment
size and greater circularity) and another is structurally dependent (i.e., results in
larger fragment size and lower circularity). Note that the microstructure-dependent
regime appears to correlate with defect parameters (i.e., spacing), and the
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 11

structurally dependent regime corresponds to the coalescence of the axial and


transverse cracks. This information is then associated with the defect parameters
(orientation, size, and spacing) and compared with theoretical predictions of
average fragmentation size distributions with respect to strain rates. The
identification of the two unique regions of fragmentation for boron carbide and the
correlation between fragmentation and defects provide important insights (e.g.,
characteristic length scale) for the fragmentation modeling of boron carbide and
materials design.

Figure 6: Optical microscope image of the boron carbide microstructure illustrating: (a)
The various types of microstructure defects and inclusions; (b) converted monochrome image
used to determine the spacing between graphite disks; (c) and (d) plot of circularity (2(πA)0.5/P)
against major axis size for: (c) dynamic uniaxial compression and (d) dynamic biaxially
confined compression (confining pressure 500 MPa). Different fragmentation regions are
hypothesized. These images are taken from (Hogan et al. 2015).

2.1.2 The Dynamic Indentation Testing – a Modified SHPB System

The dynamic indentation technique is a modified version of the SHPB system


that it is used to probe the material responses at even higher pressure under dynamic
loading. The expression of the pressure, P, induced directly under the indenter tip
is extracted from the expanding cavity model (Satapathy 2001):

𝑃 2 1𝐸
= [1 + 𝑙𝑛 ( 𝑐𝑜𝑡𝛼)] (2)
𝜎𝑠 3 3 𝜎𝑠
12 Li et al.

where 𝜎𝑠 is the static yield stress, E (GPa) is the Young’s modulus, and α is the
half-included angle of the conical indenter. The typical stresses that can be applied
are between 15 GPa to 20 GPa, which is close to the Hugoniot elastic limit (HEL)
reported by (Grady 1991).
Shown in Figure 7 is a schematic of the dynamic indentation setup that was
described in (Ghosh et al. 2012a). The transmitted bar in the conventional SHPB
system is replaced by a load cell, and the specimen is sandwiched between the load
cell and an indenter. A momentum trap is placed in front of the incident bar to allow
only one compressive wave to reach the indenter tip and cause indentation, while
the tensile stress waves will retract the indenter incrementally. Details on the
description of the setup can be found in (Ghosh et al. 2012a). With numerous
studies of the quasi-static indentation behavior of boron carbide reports elsewhere
(e.g., Domnich et al. 2002; Ge et al. 2004), limited dynamic indentation
investigations on boron carbide were published (e.g., Ghosh, Subhash, Lee, et al.
2007; Ghosh, Subhash, Sudarshan, et al. 2007; Subhash, Ghosh, and Maiti 2009;
Ghosh et al. 2012b; Subhash et al. 2016; DeVries et al. 2016b). This technique is
used to study boron carbide because it is believed to generate a similar damage
pattern (i.e., similar stress state) to early stages of impact experiments, and it can
be achieved at a much lower cost with greater accessibility.

Figure 7: Schematic of the experimental setup for dynamic indentation hardness


measurements. This image is taken from (Ghosh et al. 2012a).

Some interesting results are observed when comparing the hardness values of
boron carbide to another advanced ceramic such as silicon carbide. Shown in Figure
8 is an example in (Ghosh et al. 2012b) demonstrating a reverse rate-dependent
relationship for hardness in boron carbide. While silicon carbide (SiC) presents an
increased hardness under high strain rate indentation, boron carbide presents a
slight decrease in hardness at higher strain rates. This raises concerns regarding the
potential lack of performance of boron carbide under high-velocity impact loading.
Extensive studies have been dedicated to explaining this effect in boron carbide
(e.g., Dandekar 2001; Korotaev, Pokatashkin, and Yanilkin 2016), and structural
amorphization is attributed as the reason why this behavior occurs.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 13

Figure 8: Comparison of static and dynamic hardness values (a) in a boron carbide
ceramic revealing a decrease in hardness and (b) in a SiC ceramic revealing an increase in
hardness, under high strain rate loading. These images are taken from (Ghosh et al. 2012b).

(Ghosh et al. 2012b) used the Raman spectra, shown in Figure 9, of boron
carbide fragments from uniaxial compression and indentation experiments to
identify the occurrence of amorphization under high pressure. Also included in this
figure are the results form quasi-static and dynamic uniaxial compression using an
SHPB system. In short, Raman spectroscopy detects the vibrational, rotational and
other low-frequency transitions in material molecules by using inelastic scattering
of monochromatic light. In Figure 9, the first few peaks before 1200 cm-1
correspond to the characteristic peaks of boron carbide. Then, any additional peaks
indicate phase changes in the material. It is observed that no additional peaks other
than the characteristic boron carbide peaks occur from the uniaxial compression at
both strain rates, however, three extra peaks (named “D-peak,” “G-peak,” and
“unknown”) occur from both static and dynamic indentation. The “D-peak” is
commonly attributed as the amorphous peak in boron carbide, and several studies
have been conducted focusing on measuring this peak (e.g., Ghosh, Subhash, Lee,
et al. 2007; Ghosh, Subhash, Sudarshan, et al. 2007; Subhash et al. 2013) as a
method to quantify amorphization behavior. This also suggests that high pressure
is a necessary condition for the occurrence of the amorphous phase, but not
necessarily high loading rates. Extremely high pressures are usually generated in
most of the dynamic structural applications (e.g., impact), and thus, amorphization
is proven to be the outcome specifically for boron carbide under these conditions.
Therefore, it is crucial to account for this phase change in boron carbide material
models. Furthermore, amorphous boron carbide, as a weaker phase due to the lack
of crystalline structure, will be a preferred site for fracture initiation and
fragmentation.
14 Li et al.

Figure 9: Raman spectra collected from the static and dynamic indented regions as well as
from the surfaces of the compression fragments of boron carbide ceramic under static and
dynamic loads. For comparison purposes, a Raman spectrum from the undeformed boron
carbide is also shown from (Ghosh et al. 2012b).

In other hardness-related experiments, the damage zone of boron carbide under


the indenter tip was studied by (Ghosh, Subhash, Sudarshan, et al. 2007), and it was
correlated to the amorphization behavior. Figure 10 demonstrates the differences in
the subsurface damage zone from quasi-static and dynamic indentations. The
observations reveal significantly wider damage zones under dynamic indentation
while the depths are similar in both cases, and several cracks are seen to emanate
from the subsurface damage region in the dynamic indentation experiment. Raman
spectroscopy was conducted at these regions and an amorphous “D-peak” was
observed in the damage region under dynamic indentation, as shown in Figure 11.
On the other hand, these peaks were extremely weak in the damaged regions of the
static indentation. Overall, the greater structural damage caused by the localized
phase transformation under dynamic indentation is believed to contribute to the
reverse trend in hardness and fracture toughness. Furthermore, (Subhash, Ghosh,
and Maiti 2009) also reported a change in fracture path (i.e., from a “zig-zag”
fashion under quasi-static indentation to a “straight” path under dynamic
indentation) in another work. In a later study in 2018, (Parsard, Subhash, and
Jannotti 2018) investigated the volume change and residual stress around a
dynamically indented region of boron carbide and found a concurrent shift of the
1088 cm-1 peak (which is a characteristic peak of the crystalline boron carbide)
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 15

toward higher frequencies in the Raman spectrum. This shifting of the characteristic
peak indicates elevated compressive stresses in the crystalline regions of the
material. In addition, the loss of crystalline order in the amorphous regions is argued
to be the primary reason for volume expansion during amorphization, which in turn
causes compressive residual stress in the surrounding crystalline matrix.
Furthermore, residual stress in the material is attributed to the presence of
dislocations, stacking faults, and lattice rotations, which are likely to contribute to
further fracture in the material (J. Wang et al. 1995; Parsard, Subhash, and Jannotti
2018). These results add to our understanding of the effect of amorphization on
fracture and fragmentation. This connection is currently not well understood and
presents a challenge for current boron carbide models. The modeling of
amorphization as it relates to fracture in boron carbide will be discussed in Section
3.1.

Figure 10: Subsurface damaged region beneath (a) static indentation and (b) dynamic
indentation at a load of 19.6 N for 1.6 mm grain size boron carbide. The small arrows indicate
cracks emanating from the boundary of the damaged region. These images are taken from
(Ghosh et al. 2007).
16 Li et al.

Figure 11: Visible Raman spectra obtained from the subsurface regions of static and
dynamic indentations as well as from polished surface away from the indented region. This
image is taken from (Subhash et al. 2008).

TEM investigations have also been conducted on boron carbide to identify the
size and number of the amorphous shear band, which is considered the main
deformation mechanism and failure mode in boron carbide subjected to high-
pressure loading. Shown in Figure 12 is an example from (Ge et al. 2004)
demonstrating the formation of amorphous shear bands under a 100 mN Berkovich
indent. In general, the amorphous regions are believed to be favored sites for crack
initiation and propagation. Figure 13 shows another example of a scanning
transmission electron microscopy (STEM) investigation of the amorphous phase in
the vicinity of a residual indentation crater. The amorphous shear bands within the
deformed area are indicated and high-resolution STEM is used to examine the
lattice structure of the amorphous phase. These shear bands are potential sites for
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 17

crack initiation and propagation and capturing these microstructural changes will
be essential in boron carbide fracture modeling.

Figure 12: (a) Plane view STEM & TEM micrograph of a 100 mN Berkovich indent. (b) A
magnified image showing the amorphous bands along the (113) and (003) planes. (c) and (d)
The lattice images corresponding to the boxed area in (a) and (b). (e) A primarily amorphous
region within the Berkovich indent. These images are taken from (Ge et al. 2004).
18 Li et al.

Figure 13: Amorphous shear bands within the deformed area after indentation. (a) Low
magnification bright-field and (b) dark-field STEM images of the deformed region. The shear
bands and micro-cracks are indicated with arrowheads. Scale bar is 500nm. (c) Bright-field
STEM image of a zoomed-in shear band with a length of ~ 200nm. The black arrowhead shows
the tip of the shear band. Scale bar is 50nm. (d) ABF-STEM image of the shear band taken from
the white-box region in (c). The high-resolution STEM image shows the amorphous structure of
the shear band with a width of ~ 2nm along the [101̅1] crystallographic direction. Inset FFT
patterns demonstrate the amorphous nature of the shear band. The slight mismatch in the
orientation between the top and the bottom crystals can also be seen from the corresponding
FFT patterns. Scale bar is 2 nm. These images are taken from (Reddy et al. 2013).

Other modifications have been made to the conventional SHPB system to


access different loading conditions, including shear, tension, bending, and
intermediate strain rates. For example, (Leong et al. 2018) performed dynamic
three-point bending tests coupled with x-ray phase-contrast imaging on boron
carbide to assess the internal fracture up to ~ 800 s-1 strain rate (see Figure 14).
Unlike in high-speed imaging which can only visualize surface deformation, x-ray
phase-contrast imaging is capable of visualizing internal fracture features. This is
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 19

important because the internal fracture behavior is known to be vastly different


from surface fracture due to the heterogeneous stress state at the crack tip
(Bouchaud et al. 2012).

Figure 14: (a) A schematic of the three-bar Kolsky apparatus for high rate three-point
bending; (b) Three-point bend loading (KI = 2.4×105 MPa √𝑚s−1) of boron carbide shows
crack initiation beginning from the backside of the notch to the front. The crack tip positions
along the crack front are marked by white and black filled differently shaped markers. These
images are taken from (Leong et al. 2018).

2.2 Shock Experiments

A “shock experiment” is a general term used to refer to experiments which


generate a shock wave within the specimen. This includes plate impact experiments
(e.g., Raiser, Clifton, and Ortiz 1990; Rosenberg, Brar, and Bless 1991; Raiser et
al. 1994), laser shock experiments (e.g., Gray III 1990; Field et al. 2004; Zhao et
al. 2016), and ballistic impact experiments (e.g., Anderson and Morris 1992;
Woodward et al. 1994; Anderson and Royal-Timmons 1997). Typical strain rates
in these experiments are between 104 s-1 and 107 s-1. Shock experiments are used to
determine the Hugoniot elastic limit and equation of states of materials, as well as
to study the fracture and fragmentation of the material under extremely high
pressures and loading rates. The complex stress states generated in a shock
experiment are close to real-life impact events, where the damage mechanisms
observed in the experiments can guide model development and validation. Due to
the relatively high cost and difficulties associated with such experiments, limited
20 Li et al.

literature exists on the shock responses of boron carbide (e.g., Gust and Royce
1971; Dandekar 2001; Vogler, Reinhart, and Chhabildas 2004; Zhang et al. 2006).

2.2.1 Plate Impact Experiments

Plate impact experiments have been conducted on boron carbide by (Grady


1994; Vogler, Reinhart, and Chhabildas 2004). Shown in Figure 15 is a schematic
of the plate impact experiment that was described in (Vogler, Reinhart, and
Chhabildas 2004). The dimensions of a plate impact setup are chosen to assure that
a uniaxial strain condition is met at the central region of the specimen. In the work
of (Vogler, Reinhart, and Chhabildas 2004), the projectile velocity was measured
using three electrical self-shorting pins to an accuracy greater than 0.5%, while
additional pins were used to measure impact planarity. The projectile velocities in
(Vogler, Reinhart, and Chhabildas 2004) were between 1 km/s and 7 km/s, and the
projectile velocities in (Grady 1994) were approximately 2.5 km/s. The particle
velocity history at the target-window interface is measured using a velocity
interferometer system for any reflector (VISAR). A similar setup was used in
(Grady 1991; Grady 1994) and details of the setup can be found in their works.

Figure 15: Schematic of setup for reshock and release (plate impact) experiments. The
image is taken from (Vogler, Reinhart, and Chhabildas 2004).

(Grady 1994) compared the shock and release wave profiles between boron
carbide and silicon carbide and observed significantly different post-yield
behaviors for the two ceramics. Shown in Figure 16 is an example of the complete
shock wave profile of boron carbide and silicon carbide from (Grady 1994). Note
that boron carbide exhibits a higher HEL (~ 18 to 20 GPa). While silicon carbide
shows increasing strength beyond the initial dynamic yield (post-yield), boron
carbide demonstrates a dramatic loss in strength-supporting capability.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 21

Furthermore, the release path of boron carbide is closely related to the calculated
hydrodynamic behavior, suggesting near granular flow behavior with sustained loss
of strength. Altogether, the stress wave profiles suggest a heterogeneous
deformation in boron carbide under shock loading in contrast to the homogeneous
deformation in silicon carbide. A scatter plot of the historical data obtained for the
Hugoniot data and the estimation of shear stress and strength of boron carbide was
presented in (Vogler, Reinhart, and Chhabildas 2004) and they are shown here in
Figure 17 and Figure 18, respectively. The details in computing the HEL and the
post-yield strength using the stress wave profile can be referred to in (Vogler,
Reinhart, and Chhabildas 2004). Similar data has also been presented in (Zhang et
al. 2006) using plate impact testing.

Figure 16: Shock and release wave profiles for silicon carbide and boron carbide
ceramics measured with velocity interferometry diagnostics. This image is taken from (Grady
1994).
22 Li et al.

Figure 17: Stress–volume data from (Vogler, Reinhart, and Chhabildas 2004) and other
researchers. This image is taken from (Vogler, Reinhart, and Chhabildas 2004).

Figure 18: Shear stress 𝜏ℎ and strength Y of boron carbide in the shocked state estimated
from reshock and release experiments. This image is taken from (Vogler, Reinhart, and
Chhabildas 2004).

(Vogler, Reinhart, and Chhabildas 2004) discussed the possibilities and scale
of phase transformation (i.e., an amorphous phase), which was proposed to be
responsible for the apparent decrease in post-yield strength under shock loading
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 23

from the stress wave standpoint. They concluded that the current results are not
conclusive regarding the existence of one or more phase transitions but highly
suggestive of their existence. Note that the amorphous phase has been shown to
exist in post-shocked boron carbide, but their size and the fraction of the material
being transformed, and hence, their effect on the overall performance of the
material is still not entirely conclusive. In addition, the plate impact technique can
also be used to determine the spall strength of boron carbide (e.g., Rajendran and
Grove 1996; Dandekar 2001; Bourne 2002; Paris et al. 2010). (Paris et al. 2010)
documented a detailed review of the studies on spallation of boron carbide and
calculation of the spall strength, from which the spall strength of boron carbide was
reported between 0 and 1.2 GPa with estimated peak stress between 0 and 18 GPa
(impact velocity between 0.052 and 0.900 km/s). Note that the spall strength
decreases with increasing peak stress (impact velocity).

2.2.2 Laser Shock Experiments

Only very recently have laser shock experiments been performed on boron
carbide because of the high cost, low accessibility, and novelty of the technique.
(Zhao et al. 2016) conducted the first laser shock experiment on boron carbide, and
this is the only published laser shock result for boron carbide in the literature at the
time of this publication. Shown in Figure 19 is a schematic of the laser shock-
recovery and the velocity interferometer system for any reflector setups. VISAR is
a time-resolved velocity measurement system that uses laser interferometry to
measure the surface velocity of solids moving at high speeds (Dolan 2006). A
detailed description of the setup can be referred to in (Zhao et al. 2016). Overall,
some advantages of the laser shock technique are: 1. Enable boron carbide to be
shock compressed under controlled and prescribed uniaxial strain loading
conditions; 2. Ensure the integrity of the specimen by sending a nanosecond
duration high energy pulse which is faster than the characteristic time for crack
propagation (in microseconds), following which the recovered specimen can be
characterized post-shock by TEM; 3. Ensure sufficient shock pressure is provided
(45 to 50 GPa) which is well above the believed amorphization threshold of
approximately 20 GPa (Grady 1994).
24 Li et al.

Figure 19: Laser shock-recovery assembly, free-surface VISAR, and determination of


shock parameters. (a) Schematic drawings of the shock-recovery experiment and (b) VISAR
experiment. (c) Temporally resolved VISAR fringes, showing the shock break-out and pull-back
features. (d) Free-surface velocity, Ufs vs. t profiles. Two independent VISAR channels with
distinctive Etalon length were used to unambiguously determine the free-surface velocity. Peak
Ufs ∼ 4.2 km/s, rendering Up ∼ 1/2 Ufs = 2.1 km/s. (e) Determination of shock stress by
impedance match shock Hugoniot (σs vs. Up) of aluminum and boron carbide. These images are
taken from (Zhao et al. 2016).

“Directional amorphization” of boron carbide was introduced by (Zhao et al.


2016) in their study, which refers to the orientation of amorphous bands along the
maximum shear direction. Figure 20 demonstrates the crack initiation and
propagation from the shocked surface. Planar faults are observed in the vicinity of
the angled crack, where high-resolution TEM identifies the lattice distortion and
deviatoric strain field near the planar faults. Figure 21 shows a high-resolution TEM
image where a directional amorphous band is observed far away from the crack.
Lattice rotation is observed at the band tip, and a much higher shear strain
component is observed along the band.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 25

Figure 20: TEM/HRTEM micrographs of recovered boron carbide from laser shock
compression. (a) Low-magnification TEM image shows the shocked surface with a crack in the
center. (b) Planar fault can be identified and HRTEM image shows successive lattice
disregistries (marked by red triangles) lying along the interface. (c) The Fourier filtered image
of the boxed region in B, using (101) and (1̅01̅) reflections. (d) Corresponding geometric
phase analysis shows the deviatoric strain field in the vicinity of the planar fault. These images
are taken from (Zhao et al. 2016).

Figure 21: HRTEM micrographs of an amorphous band far away from the crack. (a) Both
ends of the amorphous band, which exhibits an ellipsoidal shape, terminate in material (one
end is shown here). (b) Lattice image at the tip of the amorphous band shows clear lattice
rotation. (c) Lattice image showing the amorphous region (marked as a-B4C) with inset
showing the corresponding FFT diffractogram. (d) Geometrical phase analysis corresponding
to (c) shows that the local shear strain is significantly higher in the amorphous region than its
surroundings, indicating that shear stress plays a crucial role in amorphization. These images
are taken from (Zhao et al. 2016).

Important results from the observation of directional amorphization behavior


of boron carbide from the laser shock experiment include: 1. It suggests that shear
26 Li et al.

stress is critical to the amorphization process; 2. The formation of amorphous bands


leads to large magnitudes of atomic distortion, which can cause further plastic
deformation, such as dislocation and twinning; 3. A transition of failure mode from
crack nucleation/propagation to amorphization is proposed at extremely high strain
rates, and a shock threshold between 25 GPa and 50 GPa has been established for
the onset of amorphization; 4. Amorphization, as a highly localized process for
atomic rearrangement, could act as an additional fast energy dissipation mechanism
during high strain rate impacts.

2.2.3 Ballistic Impact and Long Rod Penetration Experiments

As mentioned previously, boron carbide has been considered as one of the


preferred materials for light-weight body armors. Numerous works have been
dedicated to studying the fracture and fragmentation behaviors of boron carbide
during ballistic impact (Orphal et al. 1997; Westerling, Lundberg, and Lundberg
2001; LaSalvia, Leavy, et al. 2009; Savio et al. 2011). Throughout these studies,
some interesting impact responses have been identified on boron carbide, which
were also summarized in (Hogan et al. 2017): 1. The fragmentation behavior has
been observed to change beyond an impact velocity of 850 m/s (Moynihan,
LaSalvia, and Burkins 2002); 2. The velocity of the damage wave (i.e., velocity of
the primary, coherent fracture zone when no isolated crack centers can be
distinguished) has been shown to be constant before an impact velocity of 650 m/s,
after which it steadily increases (McCauley et al. 2013); 3. Boron carbide has been
observed to have no mescal zone (LaSalvia et al. 2007); 4. A sudden “drop” in
strength after shock (bullet impact introduces a shock loading), which hinders the
multi-hit capability of the boron carbide (Grady 1994).
Amorphization, which is associated with the strength drop after the HEL from
plate impact experiments, is also believed to be related to the change in
fragmentation behavior of boron carbide (Moynihan, LaSalvia, and Burkins 2002;
Chen, Hemker, and McCauley 2003). In one study by the authors, (Hogan et al.
2017) conducted ballistic impact experiments using 6.35 mm diameter WC
spherical projectiles at 275 m/s and 930 m/s (below and above the 850 m/s
threshold). Two important results are discussed in this study: 1. Similar to the
uniaxial compression experiments done using SHPB (see Section 2.1.1), two
distinct fragmentation mechanisms were identified for a ballistic impact event. One
is the “microstructure-dependent” regime, which forms smaller fragments that are
associated with the coalescence of fractures originating from carbonaceous defects
in the pressure-assisted densification (PAD) B4C boron carbide material, and the
other is the “structurally dependent” regime, which is believed to be associated with
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 27

structural failure (e.g., radial and circumferential cracking); 2. The “blocky” and
“shard” fragment shapes are a consequence of impact failure which correspond to
different fracture mechanisms parallel and perpendicular to the impact direction,
respectively. Here “blocky” refers to low aspect-ratio fragments while “shard”
refers to high-aspect ratio fragments. No amorphization is observed at these impact
velocities using Raman spectroscopy, thus brittle fracture and fragmentation are
thought to be the dominant failure mechanisms. Critical results are shown in Figure
22 and Figure 23, including optical microscopy investigations of the as-received
microstructures quantifying defect parameters, statistical fragment analysis, and
SEM investigations of the fracture mechanisms. Overall, the authors believed that
one can control fragment size and shape by controlling the carbonaceous defects
population in boron carbide, which will influence the ballistic performance of the
material (Krell and Strassburger 2014).

Figure 22: (top left) Schematic of boron carbide tile with hot-pressing direction labeled
and conceptual graphite disk defects. Optical microscope images of boron carbide
microstructure for (a) Through-thickness direction and (b) In-plane direction. Labeled in these
images are microstructural features (defined in the top left of image (a)) and the impact
direction of the spherical projectile (the dark circular object); (c) Scatter plot of aspect ratio vs
fragment size at 930 m/s. These images are taken from (Hogan et al. 2017).
28 Li et al.

Figure 23: (a) Optical microscope image of fragments from 930 m/s ballistic experiment
showing shard-like fragment (blue box) and blocky fragment (red box), and scanning electron
microscope images of: (b) fracture surface of blocky fragment with graphitic disks labeled, (c)
shard fragment with noticeable graphitic defects on the surface, and (d) magnified image of
shard fracture surface with graphitic disks labeled. As a reference, the as-received tile with the
hot-pressing and impact direction are labeled. Blue is used to denote images taken on planes
more-or-less normal to the hot-pressing direction, and the red used to denote images taken on
planes more-or-less parallel to the hot-pressing direction. These images are taken from (Hogan
et al. 2017).

In an earlier work, (Chen, Hemker, and McCauley 2003) studied deformation


mechanisms from fragments collected after ballistic impact experiments on hot-
pressed boron carbide tiles of (Moynihan, LaSalvia, and Burkins 2002). In those
tests, they used an armor-piercing projectile at velocities between 750 m/s and 1000
m/s. Even though impact velocities are similar to what has been used in (Hogan et
al. 2017), they observed nanoscale intragranular amorphous bands by examining
selected fragments using TEM. Apparent cleavage fracture surfaces were observed
on the fragments which were associated with the amorphous bands (see Figure 24),
and this was believed to be the cause of the decrease in ballistic performance in
high impact rates and pressures. The authors also proposed a threshold for the
activation of amorphization at 23.3 GPa (907 m/s) which is slightly lower than the
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 29

one proposed by (Zhao et al. 2016) using laser shock experiments (see Section
2.2.2), albeit for a presumed different material.

Figure 24: SEM and TEM observations of boron carbide fragments produced with an
impact velocity of 907 m/s (23.3 GPa). These micrographs illustrate (a) The irregular
morphology of the fragments; (b) The importance of the cleavage-like failure during high-rate
comminution; (c) A planar defect that emanates from a corner of the fracture surface; (d) High-
resolution image of the band and surrounding region in (c) indicating localized amorphization;
(e) High-resolution image taken at a region closer to the tip of the band showing the amorphous
zone. These images are taken from (Chen, Hemker, and McCauley 2003).

Other data that could be generated from a ballistic or long rod impact test
include optical microscopy visualization aimed at quantifying the damage zone
under the projectile tip (Figure 25), flash x-ray radiographs for studying the dwell
and penetration phases of the targets (Figure 26), real-time visualization of the
fracture and fragmentation behaviors of the tiles using high-speed imaging
techniques (Figure 27), and postmortem analysis on the damaged targets identifying
fracture and fragmentation mechanisms as well as damage parameters (Figure 28).
We emphasize a recent work carried out by the authors on in-situ visualization of
30 Li et al.

boron carbide fracture and fragmentation during a ballistic impact test (see Figure
27). Three high-speed cameras were used to film the front and side of the free-
standing tile. Radial cracking, circumferential damage zone, and different stages of
fragmentation patterns can be readily seen from the high-speed videos. Through
post-analysis, information such as damage velocity, projectile deceleration, and
fragment expanding velocity can be determined. All of the above information
provides valuable data for assessing boron carbide performance under ballistic
impact, as well as in material modeling and model validation.
Other impact experimental techniques such as edge-on-impact experiments
have been performed by (Strassburger 2014) on PAD boron carbide. The fracture
pattern, single crack velocity (i.e., nucleation of crack center), and fracture front
velocity (i.e., damage velocity) can be visualized and determined as a function of
impact velocities (Figure 29). It was observed that the damage velocity increased
with increasing impact velocity, and a plateau existed for impact velocity between
100 m/s and 700 m/s. Furthermore, at impact pressures close to the HEL of boron
carbide (15 – 20 GPa), a dramatic increase in damage velocity was observed. These
measurements indicate a sudden loss in damage resistance (i.e., material strength)
of boron carbide near the HEL, which correlates well with the observations of plate
impact tests (see Section 2.2.1). For other edge-on-impact examples on brittle
materials, the readers can refer to (McCauley et al. 2013). A better illustration of
the fracture front and crack center can be found in Figure 6 of (McCauley et al.
2013). Altogether, these experiments aid in the evaluation of ballistic impact
performance for boron carbide and other advanced ceramics, generate qualitative
and quantitative measurements of damage, and support the development and
validation of large-scale impact simulations.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 31

Figure 25: (a) Negative-image optical micrograph of the polished cross-section of the
impact cone region clearly showing different damage features and displaced material. (b)
Optical micrograph montage of the cone region with different features indicated. This image is
taken from (LaSalvia, Leavy, et al. 2009)
32 Li et al.

Figure 26: (a) A schematic of the modified-target flash-radiography experiments using an


APM2 bullet at ~ 850 m/s; (b) Time-resolved flash x-ray radiographs showing dwell and
penetration of the bullet into the boron carbide tile. These images are taken from (Anderson et
al. 2005).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 33

Figure 27: Ultra-high-speed imaging of ballistic impact on a boron carbide tile recently
conducted by the authors using a 30-06 M2 AP (Armor Piercing) projectile impacting at
approximately 950 m/s. (a) 5 frames of the impact from an iX camera capturing at 50,000
frames per second showing the full face of the tile. The radial and circumferential structural
damage zone can be identified in the images; (b) 5 frames of the impact from a Shimazu HPV-
X2 camera capturing at 1,000,000 frames per second showing the front of the tile. The
expansion of the debris cloud and projectile erosion can be correlated with time; (c) 5 frames of
the impact from a Shimazu HPV-X2 camera capturing at 500,000 frames per second looking at
the side of the tile. Different levels of debris cloud can be identified, as well as dwell, projectile
erosion, and tile fragmentation. Deacceleration of the projectile after penetration and debris
cloud can be measured.
34 Li et al.

Rupture Rupture Surfaces Top Surface


Surfaces
Side Surface

Side Surface Top Surface

Rupture
Face
Stagnation Point
Region

Core
Region

Rim Region

Flexural Cracks

Figure 28: Division of top surface into stagnation point, core, and rim regions of a boron
carbide tile after impacting at 1198 m/s using a long rod penetrator. Close-up of W-Ti metallic
mass which defines the stagnation point region. A partially intact truncated impact cone can be
seen. (a) Top-down view of the partially intact truncated impact cone; (b) Side-view of the cone.
The three distinct regions indicate a unique dwell and penetration mechanisms of boron
carbide compared with other advanced ceramics. SEM investigation can then be used to probe
the exact fracture mechanisms at each region. These images are taken from (LaSalvia, Leavy,
et al. 2009). New labels were added for better visualization; the text is the same as in the
original publication.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 35

2.2 μs 3.7 μs 5.2 μs

6.7 μs 8.2 μs 9.2 μs


Figure 29: A schematic of the edge-on-impact setup; (b) Path-time histories of fracture
front and cracks at 100 m/s; (c) Selection of 6 high-speed photographs from test at 100 m/s.
These images are taken from (Strassburger 2014).

2.3 Other experimental techniques that probe the fracture and


fragmentation behavior of boron carbide

2.3.1 Mechanical testing on pre-damaged boron carbide

It is generally expected that a material will behave differently when internal


damage has already been introduced to the material. Understanding the effects of
internal pre-existing cracks on the mechanical properties and fracture and
fragmentation behaviors will help us gain insights into the process of damage
evolution during impact and, eventually, allow us to design materials that are more
impact resistant. (Chocron et al. 2012) first performed quasi-static confined
compression experiments on thermally-shocked boron carbide, and the extent of
damage was characterized by sectioning the specimens. Figure 30 summarizes the
important results from (Chocron et al. 2012) including the stress-strain curves of
damaged boron carbide (see Figure 30-a), comparison of strength between intact
and damaged boron carbide (see Figure 30-b), and confinement effect on the
fragmentation behavior of boron carbide (see Figure 30-c). Figure 30-a shows that
36 Li et al.

some “residual strength” remains in the confined specimen after the maximum
compressive strength is reached. Figure 30-b shows a strength vs. confinement
pressure plot, where the degradation of strength due to the pre-existing damage is
readily seen at similar confinement pressures. Furthermore, the fragmentation
behavior of boron carbide after confined compression is shown in Figure 30-c. It is
observed that less fragments are generated at higher confining pressures. This is
attributed to the suppression of internal cracking by the higher confinement, and
the material remains more “intact” under higher confinement pressure. In addition,
a slip plane is observed at the surface of the pre-damaged specimens after confined
compression, and this is attributed to the frictional sliding between two newly open
crack surfaces at a certain angle in the specimen. This has been observed in many
rock types (e.g., Griggs 1936; Donath 1961; Li et al. 2018).

Figure 30: (a) Stress-strain curves of two intact specimens (BC-02 and BC-04) and four
pre-damaged boron carbide specimens tested under quasi-static compression at different
nominal confinement pressures. (b) Strength of intact and pre-damaged boron carbide
specimens tested under different confinement pressures; (c) Fragmentation behavior of the pre-
damaged boron carbide specimens after removal of Teflon sleeve confinements and tested at
confining pressures of (a) 50 MPa, (b) 150 MPa, (c) 350 MPa, (d) 500 MPa. These images are
taken from (Chocron et al. 2012).

A more recent study conducted by (Krimsky et al. 2019) using x-ray computed
tomography (XCT) and ultra-high-speed imaging to quantify the damage network
in boron carbide and visualize the in-situ fracture and fragmentation behavior of
pre-damaged boron carbide. The damage was introduced by thermal shocking, and
the test was done under dynamic uniaxial compression using an SHPB. Shown in
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 37

Figure 31 is an example of the stress-time history curve coupled with high-speeds


video images demonstrating different fracture and fragmentation mechanisms of
the damaged boron carbide when compared with the intact specimen (see Figure 3
in Section 2.1.1). Large branching cracks can be seen to propagate through the
specimen quickly, and relatively large fragments are obtained during failure in the
damaged samples. Overall, this information is crucial in models considering
material failure (e.g., Deng and Nemat-Nasser 1992; Paliwal and Ramesh 2008).
The pre-damaged crack network may dominate the microstructural features (i.e.,
defect population) when it comes to fracture and fragmentation behavior since the
existing cracks would be more favorable for fracture. In addition, significant
strength degradation (decreases from ~ 4 GPa to ~ 0.7 GPa) is observed and can be
correlated with the extent of damage in the material. Note that the compressive
strengths of the damaged specimens presented in (Krimsky et al. 2019) aligned
reasonably well with those of (Chocron et al. 2012) at zero confining pressure,
given that similar thermal shock cycles were carried out in both studies.

Figure 31: Stress-time history of dynamic uniaxial compression of thermally damaged


boron carbide with time-resolved high-speed video images. This image is taken from (Krimsky
et al. 2019).

2.3.2 Compaction on Granular Boron Carbide

The granular flow of brittle materials, in this case boron carbide, is usually the
final state of a material after complete fracture and fragmentation in different
applications. Applications that involve granular flow include planetary science
(Ramesh et al. 2015), mine blasting and mining exploration (Terzaghi, Peck, and
Mesri 1996; Renzo, Paolo, and Maio 2004), and terminal ballistics (Shockey et al.
1990). As mentioned by (Meyer et al. 1997) and later by (Anderson Jr 2009) (Figure
38 Li et al.

30-a in Section 2.3.1), there is a significant amount of “residual strength” that exists
in the granular state, where the granular material will have an effect on the later
impact events (e.g., further projectile erosion after bullet impacting the target).
Despite the importance of granular behavior on ballistic performance (Krell
and Strassburger 2014), limited works have been done to study the granular
behavior of boron carbide. Figure 32 shows an example on the quasi-static confined
compaction of boron carbide particles which was recently performed in the authors’
research group (journal article is under review at the date of this chapter). The
hydrostatic pressure vs. porosity curves of boron carbide (see Figure 32-a) shows a
trend where the crushed porosity increases with increasing particle size, and
cleavage steps and shallow surfaces are observed (see Figure 32-b). This
observation agrees with previous SEM/TEM studies on boron carbide fragments
from SHPB and ballistic impact experiments (see Figure 5 in Section 2.1.1 and
Figure 24 in Section 2.2.3), where transgranular fracture resulting in “smooth”
fracture surfaces is identified. Figure 32-c shows distinct fragment distributions and
non-consistent aspect ratios of boron carbide particles after compaction. For boron
carbide, two distinct regions of fragment size are observed, and the smaller
fragments preserve a flake-like shape rather than the more circular shape for the big
fragments. This agrees well with observations for the fragmentation behavior of
boron carbide by ( Hogan et al. 2015; Hogan et al. 2017) (see Figure 6 in Section
2.1.1 and Figure 22 in Section 2.2.3).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 39

Figure 32: (a) Hydrostatic pressure vs. Porosity curves showing a reverse trend of boron
carbide when compared to alumina; (b) SEM micrograph showing fracture mechanisms of boron
carbide particles under compaction; (c) Fragments of boron carbide particles under compaction
showing distinct fragment size distributions and non-consistent aspect ratios.

2.3.3 Diamond Anvil Cell (DAC) Experiment

The Diamond anvil cell (DAC) experiments coupled with Raman spectroscopy
can also be performed on boron carbide to probe amorphization under high
pressurization and depressurization. One advantage of the DAC setup is the
controlled process where in-situ scans of the Raman shift can be performed. (Yan
et al. 2009) studied the amorphization of boron carbide under a depressurization
process and concluded that the amorphous phase occurs at a critical pressure
between 13 GPa and 16 GPa (see Figure 33). This result further enhances our
understanding of the amorphization phenomenon in boron carbide. This is
important, again, because amorphization is believed to be the cause of material
fracture and fragmentation at some extreme high loading and loading rates.
Depressurization-induced amorphization may be a new strain energy/elastic energy
release mechanism that can be built into the models to incorporate amorphization-
induced fracture and fragmentation (e.g., Betranhandy, Vast, and Sjakste 2012;
Clayton 2012, 2015).
40 Li et al.

Figure 33: High-pressure Raman spectra of single-crystal boron carbide loaded in DAC
with boron carbide powder as a pressure transmitting media (PTM). (a) Raman spectra of
boron carbide loaded from ambient up to 50 GPa, (b) Unloaded from 50 GPa to ~ 1.4 GPa;
and (c) Raman spectra of recovered boron carbide pressurized to 25.9, 35.8, and 50 GPa. (d)
Optical micrograph of the recovered sample pressurized to 50 GPa. Cracks and surface relief
reveal that the crystal is loaded under a non-hydrostatic state. The Raman spectra shown in (c)
are acquired from the region free of cracks, for example, the marked region in (d). These
images are taken from (Yan et al. 2009).

2.4 Section Remarks

In this section, we summarized experimental techniques which have been used


to study fracture and fragmentation of boron carbide. These techniques incorporate
various stress states and strain rates, and the data generated can be implemented
into fracture and fragmentation models (which will be discussed in the next
section). Different initial material states (i.e., intact and pre-damaged) are also
discussed in conjunction with these experimental techniques and have shown
significantly different fracture and fragmentation behaviors. In summary,
transgranular and cleavage fracture (LaSalvia, Normandia, et al. 2009), wing crack
formation (Farbaniec, Hogan, and Ramesh 2015), shear-induced amorphous bands
and amorphization-induced cracking (Reddy et al. 2013), microstructure-dependent
and rate-dependent fragments size distribution (Hogan et al. 2015), and the unique
damage zone under ballistic impact (LaSalvia, Leavy, et al. 2009) are some
characteristic fracture and fragmentation mechanisms which have been identified
for boron carbide. These mechanisms eventually manifest as unique trends in the
mechanical properties of boron carbide (i.e., reverse rate-dependent hardness
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 41

(Ghosh et al. 2012a), strength drop after HEL (Grady 1994), and ballistic response
(Hogan et al. 2017). Altogether, they have profound effects on the material
performance, and in-depth understanding these mechanisms will allow us to better
design new generation of advanced ceramic systems.

3. MODELING FRACTURE AND FRAGMENTATION OF


BRITTLE CERAMICS AND BORON CARBIDE
In this section, models which are used to capture the fracture and fragmentation
behavior of advanced ceramics and boron carbide will be presented with theories
and simulation results. Specifically, three categories of models including micro-
mechanical models, impact/shock models, and fragmentation models will be
reviewed. These models either capture some specific features associated with
fracture and fragmentation in boron carbide, or simulate the critical outcomes of
some loading events that result in fracture and fragmentation. Note that the
variables used in the equations in the following sections may be repeated, but all
repeated variables are re-defined for clarity each time they are used.

3.1 Micro-mechanical Modeling of Brittle Ceramics

Micro-mechanical models attempt to predict the behavior of materials by


incorporating the physical mechanisms that are observed during failure. For
advanced ceramics, these mechanisms could include micro-cracking (Arrowood
and Lankford 1987), dislocation-based plasticity (Lankford et al. 1998) and
amorphization (Chen, Hemker, and McCauley 2003). In this physics-based
approach, the development of micro-scale phenomena is simulated to predict the
macro-scale response and failure mode of the bulk material, including axial
splitting, localized shear, and dilatancy. Modeling of brittle failure is further
complicated by the fact that the activation of failure mechanisms is stress state and
strain rate dependent (Ramesh et al. 2015). Due to the complexity of the brittle
failure process, simplifying assumptions are often made to make the problem more
tractable in micro-mechanical models. Common assumptions include a dilute
concentration of initial flaws, in which case crack interaction effects are negligible
(Budiansky and O’Connell 1976); plane strain or plane stress, in which case the
problem is treated in two-dimensions (Nemat-Nasser and Horii 1982); and identical
flaw sizes (Horii and Nemat-Nasser 1985). However, while these assumptions may
simplify the problem, they also limit the range of conditions for which the models
are valid. Advancement in micro-mechanical models are characterized by the
42 Li et al.

incorporation of new mechanisms as well as a stripping away of previous


simplifying assumptions to present a more realistic formulation of material physics.
This section reviews the development of micro-mechanical models for brittle
failure under compression, highlighting the physical mechanisms that have been
incorporated for different loading conditions.

3.1.1 Modeling Micro-crack Behavior in Brittle Ceramics

The nucleation, growth, and coalescence of micro-cracks is thought to be


responsible for the axial splitting observed in the failure of brittle materials under
uniaxial compression (Tapponnier and Brace 1976). Under a global state of
compressive stress, tensile micro-cracks, commonly referred to as wing cracks,
nucleate due to the stress concentration at microstructural inhomogeneities, such as
pores, inclusions, and pre-existing micro-cracks. Regardless of the initial
orientation of nucleation, experiments have shown that wing cracks grow
preferentially in the direction of axial compression. Eventually, the axial wing
cracks coalesce and cause failure by the growth of large axial cracks. (Brace and
Bombolakis 1963) proposed a sliding crack model based on the idea that frictional
sliding of pre-existing cracks cause tensile wing cracks to nucleate and grow at the
tips of the pre-existing cracks. (Nemat-Nasser and Horii 1982) later developed an
analytical solution for the sliding crack model focusing on the growth of wing
cracks under quasi-static compression. Shown on the left in Figure 34 is an example
of compression-induced wing crack growth for an arbitrarily angled pre-existing
crack, and shown on the right is a schematic of the sliding crack model used by
(Nemat-Nasser and Horii 1982).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 43

Figure 34: (Left) Experimental observation of wing crack deformation in Columbia Resin
CR 39; (Right) Schematic of sliding crack model. The pre-existing crack is denoted by PP’ and
the compression induced wing cracks are denoted by PQ. These images are taken from (Nemat-
Nasser and Horii 1982).

Key results from this model include the prediction of stable crack growth under
uniaxial and biaxial compression and unstable crack growth under lateral tension.
(Horii and Nemat-Nasser 1985) updated their wing crack model with a focus on
predicting the macroscopic shear failure that is observed in brittle materials under
moderate confinement. An array of equally spaced and identical flaws was used to
represent a suitably oriented set of flaws that would interact to create a shear fault
plane. Around the same time, (Sammis and Ashby 1986) modeled the interaction
and coalescence of micro-cracks by approximating the solid regions between cracks
as vertical columns. Aside from predicting crack propagation, models were also
developed to predict the macroscopic properties of cracked solids. (Horii and
Nemat-Nasser 1983) modeled the effective stiffness of a solid with a random
distribution of penny shaped cracks. Their results were found to agree with those
of (Budiansky and O’Connell 1976) for open cracks. However, crack closure
considerations showed that the stiffness is loading-path dependent and highly
anisotropic. Later, (Nemat-Nasser and Obata 1988) proposed a model for predicting
the dilatancy of brittle materials based on the inelastic strains associated with
frictional sliding and wing crack growth. Their model assumes that the strains
related to crack sliding and wing crack growth can be superposed on the linear
elastic strains of the matrix to obtain the effective stiffness 𝝐. The compliance tensor
for the matrix is defined as 𝑺, the Poisson’s ratio is 𝜇, and the shear modulus is 𝜇.
For a set of 𝑁 identical cracks each with an initial length of 2𝑐 and a crack normal
44 Li et al.

that is oriented at an angle 𝜃 relative to the loading direction, the crack density 𝑓 is
defined as:

𝑓 = 𝑐2𝑁 (3)

The compliance tensor for the matrix 𝑺 and the far-field stress vector is 𝝈 are used
to calculate the linear elastic contribution to the overall strain. Displacement due to
crack sliding is represented by 𝑏, and 𝑑 represents the displacement normal to the
crack face caused by crack opening. Note that both displacements have been
normalized by the crack half-width 𝑐 in this model. To analyze the growth of wing
cracks, the length of the wing crack 𝑙, the Poisson’s ratio 𝜐, and shear modulus 𝜇
are needed. Finally, the effective compliance of the solid is given by:

𝝐 = 𝑺: 𝝈 + 2𝑓𝑏𝒑0 + 2𝑓𝑑𝒑1 + 2𝑏𝑙𝒒0 + 𝑓𝑑𝑙𝒒1


1−𝜈 2 (4)
+𝑓 𝜋𝑙 [(4𝝈: 𝜶)𝜶 + (𝝈: 𝜷)𝜷]
8𝜇

where 𝒑0 , 𝒑1 , 𝒒0 , 𝒒1 , 𝜶, and 𝜷 are directional tensors used to account for the


orientations of the pre-existing crack and wing cracks with respect to the loading
direction. The first term represents the linear elastic contribution from the matrix;
the second term accounts for the strain caused by crack sliding; the third term
accounts for the strain due to opening of the pre-existing crack; the fourth term
accounts for the normal concentrated gap at the pre-existing crack tip caused by
frictional sliding; the fifth term accounts for the tangential concentrated gap at the
pre-existing crack tip caused by the opening of the pre-existing crack; and the sixth
term accounts for the strain caused by the opening of the wing cracks. Their results
highlighted the effects of friction, as it relates to crack sliding, and fracture
toughness, as it relates to crack opening, on volumetric strain. (Ashby and Sammis
1990) refined the wing crack model to relate crack extension to macroscopic stress
while accounting for crack growth and interactions. Their model was used to
generate failure surfaces for different types of rocks (e.g., sandstone).
In subsequent works, the wing crack model was further modified to account for
the effects of dynamic loading. When loaded past a transition strain rate, many
brittle materials exhibit increased strength, and modeling this rate-sensitivity has
been the focus of many theoretical works. (Deng and Nemat-Nasser 1992) extended
the two-dimensional wing crack model to dynamic strain rates. A key addition was
the calculation of the relationship between the static and dynamic stress intensity
factors using an approximate function of the crack tip speed established by (Freund
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 45

1973). The dynamic stress intensity factor K1d can be related to the quasi-static
stress intensity factor K1s as follows:

𝐾𝐼𝑑 = 𝑘(𝑙 ̇)𝐾𝐼𝑠 (5)

The relationship between the dynamic and static stress intensity factors, k(𝑙̇), was
first established by (Freund 1973) as a function of crack speed and was
approximated by (Deng and Nemat-Nasser 1992) as:

−1
𝑙̇ 𝑙̇ 𝑐̅𝑅 𝑡 ∗
𝑘(𝑙)̇ = (1 − ) (1 − ) , 𝑐𝑅 = (6)
𝑐𝑅 2𝑐𝑅 𝑐

where 𝑐̅𝑅 is the Rayleigh wave speed, 𝑐𝑅 is the number of half-flaw lengths that the
Rayleigh waves would travel during a time period of 𝑡 ∗, 𝑙̇ is the crack extension
rate, and 𝑐 is the half-flaw size. This relation has since been used by (Ravichandran
and Subhash 1995) and (Paliwal and Ramesh 2008) to account for dynamic crack
growth. Following a similar approach to (Horii and Nemat-Nasser 1985), crack
interaction effects during dynamic loading were explored by (Deng and Nemat-
Nasser 1994) using an array of equally spaced and identical flaws. Building on the
wing crack model by (Horii and Nemat-Nasser 1985), (Ravichandran and Subhash
1995) developed a micro-mechanical model for high strain rate compression of
ceramics and validated it against compression data for aluminum nitride. They
identified the relationship between toughness and crack velocity as the dominant
parameter that governs rate sensitivity in brittle materials. Later, (Paliwal and
Ramesh 2008) considered interactions between wing cracks of the most favorable
flaw orientation by a crack-matrix-effective-medium approach. In addition, rather
than assuming all flaws are of the same size, their model accounted for different
flaw size distributions and explored their effects on the dynamic mechanical
response. Shown in Figure 35 are model predictions from (Paliwal and Ramesh
2008) for the stress-strain behavior and damage evolution of a brittle material at
different strain rates. The y-axis on the left is the stress in the material, y-axis on
the right is the damage as quantified by crack density, and the x-axis is the strain.
From this plot, it can be seen that peak strength becomes more rate-sensitive after
104 s-1, damage accumulates faster at lower strain rates, and a greater level of
damage can be sustained at higher rates.
46 Li et al.

Figure 35: Predicted stress-strain (solid) and damage evolution (dashed) curves for a
nominal brittle material at different strain rates by (Paliwal and Ramesh 2008). The inset shows
the evolution of stress and damage through time.

More recently, with increasing computational power and the development of


new modeling approaches, there has been an increased focus on three-dimensional
modeling of micro-cracking using the wing-crack formulation. (Kolari 2017)
developed a three-dimensional wing crack model based on the 3D approach
proposed by (Ashby and Sammis 1990). This model is capable of predicting micro-
crack growth, crack interaction, crack induced anisotropy, and axial splitting.
(Ayyagari, Daphalapurkar, and Ramesh 2018) proposed a three-dimensional model
for computing the effective compliance of a solid with micro-crack evolution. As
previously noted, preferential crack growth in compression leads to significant
anisotropy in the compliance of the overall material. In the work of (Ayyagari,
Daphalapurkar, and Ramesh 2018), the anisotropic compliance is modeled as a
function of flaw statistics, stress state, and crack interaction. While their model was
found to compare relatively well with Westerly granite data (Zoback and Byerlee
1975), it was limited in predicting the unloading response due to simplifying
assumptions relating to inelastic strains normal to the wing cracks, non-interacting
cracks, and a lack of flaw data.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 47

3.1.2 Modeling Plasticity and Amorphization in Brittle Ceramics

In addition to micro-cracking, other deformation mechanisms such as


dislocation-based plasticity, granular flow, pore compaction, and amorphization are
also relevant to the dynamic failure process and have been integrated into micro-
mechanical models. Dislocation-based plasticity has been observed in recovered
fragments from dynamic experiments on confined ceramics (e.g., Lankford et al.
1998; Paliwal, Ramesh, and McCauley 2006). This has motivated the development
of models that account for dislocation-based plasticity under high confinement.
(Rajendran 1994) proposed a ceramic model based on a set of microphysically-
based constitutive relationships where the total strain is decomposed into elastic,
plastic and microcracking components. The micro-cracks are activated based on a
generalized criterion and the growth is controlled by fracture mechanics. This
model was subsequently used to simulate plate impact experiment of boron carbide
tiles (Rajendran and Grove 1996). (Zuo, Disilvestro, and Richter 2010) modified
their rate-dependent damage model based on the Dominant Crack Algorithm (Zuo
et al. 2006) to include plastic deformation. Other efforts to incorporate inelastic
energy dissipation mechanisms include the model by (Deshpande and Evans 2008),
which generalized the model by (Ashby and Sammis 1990) to include plasticity at
high confining pressures and the effects of microstructural parameters such as flaw
size and density, fracture toughness, friction, crack shape, and crack growth rate.
The Deshpande and Evans 2008 damage model was later combined into the
Johnson-Holmquist-Beissel ceramic model by (Holmquist and Johnson 2012).
Building on the 2D model by (Paliwal and Ramesh 2008), (Hu et al. 2015)
developed a 3D model for brittle ceramics that incorporated dislocation-based
plasticity as a deformation mechanism in addition to micro-cracking. In their
model, flaw statistics were incorporated for both flaw size and orientation. Their
results indicate that the three-dimensional orientation distribution of flaws has a
significant impact on the mechanical response of confined ceramics. Under triaxial
compression, micro-cracking was found to be suppressed while plastic flow was
triggered. Later, (Tonge and Ramesh 2016a, 2016b) presented a model for brittle
material failure that incorporated dynamically interacting flaws, granular flow, and
pore compaction. The micro-crack model incorporated is based on a wing-cracking
mechanism where micro-cracks growth is controlled by the stress intensity factor
at the crack tips. As micro-cracks coalesce during dynamic loading, the fragmented
region can be treated as a granular material at some predescribed damage level, and
further compression of the material will crush out the porosity between the
fragments. Their investigation emphasized the role of granular flow in controlling
the collapse of the strength of ceramics.
48 Li et al.

While amorphization in boron carbide has been the subject of extensive


research, micro-mechanical models incorporating amorphization as a failure
mechanism for ceramics have been limited. (Clayton 2012) applied third-order
anisotropic nonlinear elasticity to model the onset of amorphization in boron
carbide using Born instability (Born 1942) as the activation criterion. (Clayton
2013) extended his earlier model by explicitly modeling the amorphous phase.
(Clayton and Knap 2018) employed geometrically nonlinear phase field theory to
model twinning, amorphization, and fracture in boron carbide. From model
simulations, inelastic shear strain from twinning and slippage in amorphous bands
were found to cause localized fracture and loss of macroscopic strength. Later,
(Zeng, Tonge, and Ramesh 2019) developed an amorphization model that uses a
stress-based criterion to determine the onset of amorphization. In this model,
damage to the amorphous bands is modeled by increasing shear strain, and granular
flow is activated at a critical shear damage level. This model was validated against
plate impact experiments (Vogler, Reinhart, and Chhabildas 2004) and dynamic
Vickers indentation tests by (Parsard, Subhash, and Jannotti 2018). Figure 36 shows
a comparison of model predictions against the spatial distribution of amorphization
observed in the dynamic Vickers indentation experiment. It can be seen that there
is a reasonable agreement between the boundary of experimentally observed
amorphization and a predicted amorphization volume fraction of approximately
1%, which may be difficult to measure experimentally.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 49

Figure 36: Predicted amorphization fraction contour is compared against experimentally


measured boundary for amorphization at a dynamic Vickers indentation site. This image is
taken from (Zeng, Tonge, and Ramesh 2019).

3.2 Modeling Impact/Shock Behavior of Brittle Ceramics and


Boron Carbide

In order to understand the behavior of advanced ceramics under impact loading,


numerous types of models have been developed since the 1940s. For example: 1.
Hydrodynamic theory (e.g., Birkhoff et al. 1948; Tate 1967, 1969, 1986; Anderson
Jr and Walker 1991); 2. Micro-mechanical modeling (e.g., Curran et al. 1993); 3.
Dimensional analysis (e.g., Clayton 2015, 2016); 4. Spherical and cylindrical cavity
analysis (e.g., Forrestal 1986; Forrestal and Longcope 1990; Forrestal and Tzou
1997; Guo, He, and Wen 2013); and 5. Phenomenological models (e.g., Johnson
and Holmquist 1992; Johnson and Holmquist 1994; Holmquist and Johnson 1999;
Holmquist and Johnson 2006; Holmquist and Johnson 2011). The main objective
of these models is to correlate the ballistic performance parameters (i.e., depth of
penetration, projectile deceleration and erosion speed, and cavity size) to some
material properties and target geometry (i.e., tile thickness (e.g., Zaera and
50 Li et al.

Sánchez-Gálvez 1998)). This sub-section reviews three categories of


phenomenological and analytical models, including the Johnson-Holmquist (JH)
Models, the hydrodynamic theory (the Walker-Anderson model), and the
expanding cavity model. These models have been the most widely implemented
models in recent studies on impact-induced damage of brittle ceramics. Important
formulations and critical computed results will be presented along with the
corresponding models.

3.2.1 The Johnson-Holmquist (JH) Models and their Extension

The JH models have been implemented in numerous investigations to study the


behavior of brittle ceramics under impact (e.g., Anderson and Royal-Timmons
1997; Cottrell et al. 2003; Cronin et al. 2004; Holmquist and Johnson 2005). Recent
literature published on large-scale impact simulation of boron carbide incorporated
the JH models to study fracture and fragmentation behavior of the material (e.g.,
Westerling, Lundberg, and Lundberg 2001; Shokrieh and Javadpour 2008;
Krishnan et al. 2010; Fountzoulas and LaSalvia 2013; Arslan and Güneş 2017). The
first JH model was developed in (Johnson and Holmquist 1992) (often referred to
as the JH-1 model) with a later improved version in (Johnson and Holmquist 1994)
(often referred to as the JH-2 model). A more refined version of the JH-2 model,
proposed in (Holmquist and Johnson 1999), was dedicated to predicting the
response of boron carbide under large strains, high strain rates, and high pressure.
Aspects of the JH-2 model is summarized in Figure 37, where strength, pressure,
and different levels of damage are three components taken into consideration in the
model. The JH-2 model includes specific representations of strength for intact and
fractured material, a pressure-volume function that involves the bulking effect, and
a damage function that transitions the material from intact to fractured state.
Constitutive equations incorporating these components are also shown in Figure 37.
The JH-2 model was used by many authors to perform finite element simulations
of a ballistic impact on a ceramic (e.g., Anderson Jr, Walker, and Lankford 1995;
Cicchetti et al. 2014; Wang and Li 2015; Gour, Serjouei, and Sridhar 2017;
Toussaint and Polyzois 2019).
In an effort to predict the response of boron carbide under plate impact and
high-velocity ballistic impact conditions, the baseline constants of the JH-2 model
need to be extrapolated from the limited existing data of boron carbide. After a
reasonable fit for both strength and pressure, as well as assuming some initial
damage parameters (i.e., D1 = 0.001, extremely small plastic strain for brittle
material), the model has been implemented into a series of impact simulations for
further validation (Holmquist and Johnson 2006). Shown in Figure 38 is a
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 51

comparison between the computed results using the JH-2 model and test results
extracted from different literature. Figure 38 shows the comparison in plate impact
experiments by using the test data from (Grady and Moody 1996). A good
correlation between the computed and test results are obtained. A more gradual
decrease of the release wave in the test results is observed, while the model predicts
a more abrupt change. According to (Holmquist and Johnson 1999), this could be
an effect of reduced shear modulus (assumed constant in the model) with respect to
damage during material failure. Therefore, a term for damage-induced softening
shear modulus would be an improvement to the model. In the high-velocity long
rod penetration results (see Figure 39), the computed results underpredict the
penetration velocity when the baseline constants are used. They examined the
sensitivity of the model to the strength of the failed material, and the failed strength
was found to play an important role in impact predictions. It was also concluded by
the authors that underestimation of the model could mean that the actual damage
and fracture/softening mechanisms are much more complex than previously
assumed.
52 Li et al.

Figure 37: Description of the JH-2 model. This image is taken from (Johnson and
Holmquist 1994).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 53

Figure 38: Comparison of computed results using the JH-2 model and test results for four
plate impact tests of boron carbide from (Grady and Moody 1996). These images are taken
from (Holmquist and Johnson 1999).
54 Li et al.

Figure 39: Comparison of computed results by using the JH-2 model and test results for
long rod penetration tests of boron carbide. These images are taken from (Holmquist and
Johnson 1999).

(Holmquist and Johnson 2006) performed impact simulations on boron carbide


again using their latest model, which is often referred to as the Johnson-Holmquist-
Beissel (JHB) model. The improvements of the JHB model from the previous JH-
1 and JH-2 include a phase change and a slight deviation of intact and failed strength
below or after some characteristic strain rate (see Figure 40). The authors re-
examined the model parameters by using the latest experimentally available plate
impact data from (Vogler, Reinhart, and Chhabildas 2004) and provided a much
closer fit to the new plate impact data. Details on the model parameters and
computed results can be found in (Holmquist and Johnson 2006). However, some
concerns were also raised: 1. There were significant differences between plate
impact data reported by different researchers, which complicates model validation;
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 55

2. The current model parameters were only fitted for plate impact experiments, and
results were inconsistent when ballistic impact or other penetration events were
simulated. They performed another trial using the JHB model on boron carbide later
in (Johnson and Holmquist 2008) with three conditions specified: 1. Plate impact
condition (uniaxial strain, highest pressure, and stress); 2. Projectile impacting a
thick target (lower stress, higher inelastic strain); 3. Projectile impacting a thin
target (combination of compressive and tensile pressure). In the end, three sets of
individual model parameters were needed to match the experimental data. They
concluded with three perplexing observations: 1. Boron carbide appeared to be
much more brittle under plate-impact loading than under penetration loading, which
resulted in extremely small damage constants for initial model parameters; 2. There
are differences in the strength of the failed material between the plate-impact results
(higher strength at high pressures) and the results for penetration into thick plates
(lower strength at high pressures); 3. Effect of tensile strength on both the
penetration (thick plate) and perforation (thin plate) results are significant since
much of failure occur under tension. However, a wide range of tensile strength data
was reported in literature (e.g., Winkler and Stilp 1992; Grady and Moody 1996;
Dandekar 2001), which could be a consequence of different manufacturing methods
and experimental techniques. This inconsistency in tensile strength would result in
large variation in model prediction. These considerations demonstrate limitations
of the JH model, for which further research is needed.
56 Li et al.

Figure 40: Description of the JHB model. The pressure-volume on the left shows the
relationship when phase change is suppressed. On the right is the re-evaluated pressure-volume
relationship when phase change occurs. This image is taken from (Holmquist and Johnson
2006).

A recent modification to the JH model was proposed by (Dyachkov et al. 2018).


Their model modifies the damage rate equation by defining a consistent total plastic
strain component in the failed material. The number of free model parameters is
reduced, and this enables the plastic strain to be explicitly accumulated during the
failure process. In this model, the parameters are fitted with the plate impact data
obtained in (Vogler, Reinhart, and Chhabildas 2004), and the model was validated
against plate impact data obtained in (Grady 1994). As mentioned by (Holmquist
and Johnson 1999), the vast difference in plate impact data from different
researchers presents difficulties for model validation. In the model proposed by
(Dyachkov et al. 2018), this can be addressed by adjusting only the failed material
strength. Shown in Figure 41a is a typical failure path of boron carbide under plate
impact. It demonstrates the differences in damage accumulation of boron carbide
between the present study and the JH models. Figure 41-b demonstrates the fitted
curves for the intact (at HEL) and failed (after HEL) strength of boron carbide.
Shown next in Figure 42 is an example of the comparison between the improved
JH model proposed by (Dyachkov et al. 2018) and test results obtained in (Grady
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 57

1994). With only material strength adjusted based on Figure 41-b, the model
reproduces the main features of the shock velocity profiles in good agreement.

Figure 41: (a) The typical scheme for a boron carbide failure process on a diagram of
“equivalent stress-pressure” in plate-impact experiments. Path 1 used in JH-1 and JHB models
corresponds to the instantaneous failure, and path 2 used in the JH-2 model corresponds to the
gradual failure with the acceleration of the damage rate. Path 3 used in (Dyachkov et al. 2018)
corresponds to a solution of the kinetic equation with a decelerating damage rate in the
approach to the completely failed state, and path 4 shows the uniaxial unloading of a
completely damaged material; (b) The pressure dependencies of intact (at HEL) and failed
(after HEL) strength fitted for the best agreement with (Vogler, Reinhart, and Chhabildas 2004)
and (Grady 1994). These images are taken from (Dyachkov et al. 2018).

Figure 42: Simulated and experimental interface velocities for the BC4, BC5, and BC7
tests. Here the only material strength was adjusted to reproduce the experimental velocities.
These images are taken from (Dyachkov et al. 2018).
58 Li et al.

3.2.2 The Hydrodynamic Theory and The Walker-Anderson Model

The hydrodynamic theory is one of the first approaches used to model the
penetration of long projectiles into thick targets (e.g., Alekseevskii 1966; Anderson
Jr and Walker 1991; Birkhoff et al. 1948; Tate 1967, 1969, 1986). This theory
employs the Bernoulli's equation at the projectile-target interface to relate the
velocities of impact (V) and penetration (U) to the density of the projectile (𝜌𝑝 ) and
target (𝜌𝑡 ). The expression is defined as (Birkhoff et al. 1948; Pack and Evans
1951):

1 1
𝜌𝑡 𝑈 2 = 𝜌𝑝 (𝑉 − 𝑈)2 (7)
2 2

where U and V are assumed to be constant. Then, the relationship between V and U
can be established as:
𝑉
𝑈=
𝜌 (8)
1 + √𝜌𝑡
𝑝

The total depth of penetration (P), normalized by the length of the projectile (L), is
expressed as:
𝑃 𝜌𝑝
=√ (9)
𝐿 𝜌𝑡

The original hydrodynamic theory was later modified by (Tate 1967) and
(Alekseevskii 1966) to incorporate the deceleration and erosion of the projectile by
adding a projectile flow stress (𝑌𝑠 ) and a penetration resistance (𝑅𝑡 ) term to the
equation:
1 1
𝑅𝑡 + 𝜌𝑡 𝑢2 = 𝜌𝑝 (𝑣 − 𝑢)2 + 𝑌𝑠 (10)
2 2

where u and v are time-dependent (instantaneous) nose and tail velocities of the
projectile, respectively. The deceleration of the projectile (𝑣̇ ) can be expressed as:
𝜎𝑝
𝑣̇ = − (11)
𝜌𝑝 ∗ 𝑙

where 𝜎𝑝 is the dynamic compressive yield strength of the projectile, and l is the
instantaneous projectile length. In theory, the deceleration of the projectile occurs
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 59

due to the reflected elastic wave from the back of the projectile. Therefore, the
projectile erosion rate (𝑙)̇ is represented as the difference between the nose and tail
velocities:
𝑙 ̇ = −(𝑣 − 𝑢) (12)

One limitation of the hydrodynamic theory lies in determining the penetration


resistance, where a series of iterations are required in order to find good agreement
between the model and experimental data (Alekseevskii 1966). In addition, the
theory is material-dependent which adds on to the difficulties in finding parameters.
Later, (Walker and Anderson Jr 1995) proposed a time-dependent model for
long rod penetration by modifying the Bernoulli's equation approach from a
momentum balance consideration. The main constitutive equation is expressed as:

𝑧𝑖 ∞
𝛿𝑢𝑧 𝛿𝑢𝑧 1 1
𝜌𝑝 ∫ 𝑑𝑧 + 𝜌𝑡 ∫ 𝑑𝑧 + 𝜌𝑝 (𝑢2 − 𝑣 2 ) − 𝜌𝑡 𝑢2
𝑧𝑝 𝛿𝑡 𝑧𝑖 𝛿𝑡 2 2
∞ (13)
𝛿𝛿𝑥𝑧
− 2∫ 𝑑𝑧 = 0
𝑧𝑖 𝛿𝑥

where 𝑢𝑧 is the axial component of particle velocity, and 𝑧𝑖 and 𝑧𝑝 are the
instantaneous locations of the projectile nose and tail, respectively. The derivation
of the model can be found in (Walker and Anderson Jr 1995). The model was later
combined with the Drucker–Prager constitutive model and extended to ceramic
targets (Walker 2003). (Bavdekar, Subhash, and Satapathy 2019) in their recent
study identified several limitations of the Walker and Anderson model: 1. The
Drucker–Prager model parameters need to be extracted from experimental data; 2.
The non-linear behavior of ceramics shown near or above HEL cannot be accurately
captured by the bilinear Drucker–Prager model; 3. The cylindrical expanding cavity
model employed in the Walker-Anderson model consists only of a plastic and
elastic zone of damage (as opposed to the expanding cavity model proposed by
(Satapathy 2001)); 4. The inconsistency of co-ordinate systems in the model (i.e.,
Cartesian co-ordinates for momentum conservation equation, cylindrical co-
ordinates for expanding cavity model, and spherical co-ordinates for shear stress);
5. Non-physical representation of the target velocity profile; 6. The model does not
include the phase of dwell and projectile defeat. Note that the dwell phase starts
when the projectile hits a ceramic armor, followed by the tip fracture, projectile
erosion and ends when the ceramic starts to damage. Thus, it cannot be used in low-
velocity impact where no penetration occurs. These limitations lead to the
60 Li et al.

differences between the computed and test results and difficulties in model
validation.

3.2.3 The Expanding Cavity Model

(Satapathy and Bless 1996, 2000) and later (Satapathy 2001) proposed the first
modified expanding cavity model for brittle ceramics. This model derives the
damage zone of brittle ceramics caused by the sudden expansion of a cavity at a
certain pressure and defines different response regions: 1. The cavity zone; 2. The
comminuted zone; 3. The radially cracked region; 4. The elastic zone; and 5. The
undisturbed region. Figure 43 shows a comparison between the damage patterns
obtained from different impact testing and the modeled damage zone of the
expanding cavity model. This image is extracted from a review paper by (Subhash
et al. 2008). Similar damage patterns can be readily seen from different impact
experiments when extreme pressure is introduced at a relatively small area. The
expanding cavity model can, therefore, provide some qualitative measurements of
fracture in terms of damage zone and cracking.

Figure 43: Comparison of induced damage patterns from various experimental methods to
the schematic of the cavity expansion model. Top left: edge-on impact; top right: schematic of
the spherical cavity expansion model; bottom left: normal impact; and bottom right: dynamic
indentation. This image is taken from (Subhash et al. 2008).
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 61

The model can be briefly explained as follows (Satapathy 2001): 1. The impact-
induced cavity generates an elastic wave that spreads out at the material sound
speed (elastic region); 2. The hoop stress in the elastic region is tensile. Since the
tensile strength of a brittle ceramic is much lower than the compressive strength
(i.e., usually in the order of 10), radial cracks appear when the hoop stress equals
the tensile strength (radial crack zone); 3. The materials in the crack zone are
assumed to be able to only support the compressive radial stress, thus, when the
radial stress is increased to the compressive strength of the material, further
cracking occurs and leads to comminution (comminuted zone); 4. The comminuted
material is taken to behave as a Mohr-Coulomb material with zero cohesive strength
(granular flow – shear saturation). The Mohr-Coulomb model has the form:

𝐼𝑛𝑡𝑎𝑐𝑡: 𝜏 = 𝜏0 + 𝑎𝑖 ∗ 𝑃 (14)

𝐶𝑜𝑚𝑚𝑖𝑛𝑢𝑡𝑒𝑑: 𝜏 = 𝑎𝑐 ∗ 𝑃 for 𝑃 < 𝑃𝑠


(15)
𝜏 = 𝜏𝑠 for 𝑃 > 𝑃𝑠

where 𝑎𝑖 and 𝑎𝑐 are constants, 𝜏0 is the material strength in pure shear, 𝜏 is the
shear strength under multiaxial loading, 𝑃 is the hydrostatic pressure applied, and
𝑃𝑠 is the saturated pressure for shear strength. A detailed framework of the
expanding cavity model can be referred to (Satapathy 2001), which is omitted here.
(Satapathy 2001) used alumina AD995 in his study to demonstrate the
expansion velocities of cracked and comminuted fronts for different cavity
expansion velocities (see Figure 44). He showed that with the cavity expansion
velocity at around 980 m/s the cracked region disappeared (i.e., Cp = Cc), which
means the material is completely comminuted (i.e., elastic-comminuted). He then
showed the correlation of cavity pressure and penetration resistance with respect to
different penetration velocities and concluded with a reasonable agreement with
experimental data.
62 Li et al.

Figure 44: Speeds of the comminuted and cracked zones. Cc and Cp denote the speed of
crack zone and comminuted zone boundary, respectively. This image is taken from (Satapathy
2001).

(Bavdekar et al. 2017) proposed a modification to the expanding cavity model


with a focus on improving predictions for target resistance of structural ceramics.
According to the authors, the conventional Mohr-Coulomb model is a simple linear
pressure-dependent model that is not sufficient to capture the response of brittle
ceramics at high confinement pressures beyond HEL. Thus, a modified form of the
Mohr-Coulomb model is proposed with an exponential form:

∗ ∗𝑃∗
𝜏 ∗ = 𝑚 + 𝑛 ∗ 𝑒 −𝑘 (16)

where m, n, and 𝑘 ∗ are constants, 𝜏 is the shear strength and 𝑃∗ is the pressure
applied. The superscript indicates that the quantities are normalized with their
corresponding HEL values. Note that the above equation is only suitable for intact
materials. For comminuted materials, a different formula is presented:


𝜏 ∗ = 𝑎1 ∗ (1 − 𝑒 −𝑘1∗𝑃 ) (17)

𝜏𝑠
𝑎1 = (18)
𝜏𝐻𝐸𝐿

where 𝑎1 and 𝑘1 are new constants derived by assuming no shear strength in the
comminuted materials. The constant 𝑎1 controls the value at which the shear
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 63

strength of the comminuted material saturates, and hence, it can be related to the
saturated shear strength (𝜏𝑠 ). A similar form of the modified Mohr-Coulomb model
has been presented in (Shafiq and Subhash 2016a, 2016b). Figure 45 demonstrates
the good agreement between the modified Mohr-Coulomb model and the
experimental data for several ceramics. Note that the experimental data presented
in Figure 45 is for other ceramics (i.e., AlN, SiC, granite, float glass, ZrB2 -SiC),
and limited data points of boron carbide exist for shear strength (e.g., Grady 1994;
Holmquist and Johnson 2006).

Figure 45: (a) Experimental data for several ceramics illustrates the exponential
relationship between normalized pressure and shear strength. (b) Extended Mohr-Coulomb
model for intact and comminuted ceramics. 𝑇 ∗ is the normalized hydrostatic tensile strength of
the intact ceramic. These images are taken from (Bavdekar et al. 2017).

Shown in Figure 46 is the computed results for penetration resistance (𝑅𝑡 )


versus penetration velocity (U) after incorporating the modified Mohr-Coulomb
model into the improved expanding cavity model. The results are compared with
the original expanding cavity model developed by (Satapathy 2001). Much better
agreement is obtained between the experiments and the model after the
modification, especially in the case of boron carbide and aluminum nitride. Note
that the lack of experimental data hinders the optimization of the constant 𝑘1, which
will be different for each material. In this study, 𝑘1 was estimated by fitting the
current model to the available experimental data. Alternatively, a relationship
between 𝑘1 and other readily obtainable material properties will be desirable for
future development.
64 Li et al.

Figure 46: Comparison of experimental data and improved expanding cavity model results
for target resistance and penetration velocity for four structural ceramics. RMSE denotes the
root-mean-squared error between the experimental data and computed results. These images
are taken from (Bavdekar et al. 2017).

(Bavdekar, Subhash, and Satapathy 2019) proposed a unified model for dwell
and penetration during long rod impact on thick ceramic (i.e., assumes semi-infinite
in thickness) targets by combining the Walker-Anderson model (Walker and
Anderson Jr 1995) and their improved expanding cavity model (Bavdekar et al.
2017). This model is called “unified” because, unlike the previous models which
required experimental data for each material, a single set of parameters is used to
predict the response of many brittle ceramics. This model also captures the
transitional phase of dwell and penetration during long rod penetration. A detailed
derivation of the model can be referred to (Bavdekar, Subhash, and Satapathy
2019).
Shown in Figure 47-a is an example comparing the conventional hydrodynamic
model (see Section 3.2.2) and the unified model, along with long rod penetration
data of aluminum nitride impacted at 2.98 km/s. Similar to what has been defined
in the Walker-Anderson model, the depth of penetration is denoted as the difference
between the projectile nose and tail position. The Hydrodynamic model
overpredicts the nose position, and, consequently, the total depth of penetration. In
other words, it underestimates the performance of the material. The unified model,
however, appears to capture the experimental data much better. Figure 47-b shows
the loss in kinetic energy of the projectile, for which it is shown that most of the
energy loss is due to projectile erosion. A similar observation has been reported by
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 65

(Anderson and Walker 2005) in one of their efforts to model the dwell and projectile
defeat phase before penetration.

Figure 47: (a) Normalized nose and tail positions of a tungsten projectile (L/D = 20)
penetrating an AlN target compared with experimental data. Impact velocity = 2.98 km/s. The
nose position also corresponds to the instantaneous depth of penetration (p). P is the total depth
of penetration; (b) Normalized kinetic energy of a tungsten projectile (L/D = 20) penetrating an
AlN target. Impact velocity = 2.98 km/s. Most of the kinetic energy is lost due to the erosion of
the projectile. These images are taken from (Bavdekar, Subhash, and Satapathy 2019).

Figure 48 shows the goodness of fit between the models and experimental data
for four structural ceramics, including boron carbide. The Hydrodynamic model
always overpredicts both the depth of penetration and penetration, which indicates
an oversimplified consideration towards modeling the complex behavior of these
ceramics. The depth of penetration (see Figure 48-a) approaches the Hydrodynamic
prediction only at very high impact velocities. This agrees with the argument that
the Hydrodynamic model is not accurate at low impact velocities. For the
penetration velocity (see Figure 48-b), the Hydrodynamic model almost seems
parallel to the experimental data, which leads to a total overestimation. On the other
hand, the unified model (with the set of universal constants) captures the
experimental data quite well, especially those of boron carbide and aluminum
nitride. The model with the universal constant tends to underpredict the response of
silicon carbide. One possible explanation for this lack of agreement is that the
model parameters are derived from a large variety of brittle materials, where
deviation could be expected for some specific materials. Note that the universal
model parameters that the authors described have not been provided in this article,
and it is expected that some improvements are underway.
66 Li et al.

Figure 48: Comparison of experimental data and model results for (a) the normalized
depth of penetration and (b) the steady-state penetration velocity as a function of impact
velocity for a long tungsten projectile (L/D = 20) impacting different structural ceramics. These
images are taken from (Bavdekar, Subhash, and Satapathy 2019).

The Hydrodynamic-type and expanding cavity-type models described


previously constitute most of the works focused on modeling structural ceramics
analytically under impact loading. The advantages of analytical models compared
to finite element (FE) models are that they do not require iteration, and therefore,
the solutions can be obtained quickly. Unlike micro-mechanical models, which are
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 67

sometimes material-specific, analytical models are usually developed for a


representation of a certain group of materials (e.g., structural ceramics) and loading
conditions (e.g., target thickness, projectile type). With the advancement in
experimental techniques such as in-situ techniques for damage visualization (e.g.,
flash x-ray) and quantitative measurement (e.g., Photonic Doppler Velocimetry),
the model parameters could be more accurately determined. Finally, we re-
emphasize that the previously mentioned impact and shock models have been
developed for advanced ceramics in general, however, boron carbide experimental
data has been used for model validation and show reasonable agreement with the
models. This suggests that the impact and shock behavior of boron carbide can be
reasonably captured by these models.

3.3 Modeling the Fragmentation Behavior of Brittle Ceramics and


Boron Carbide

Fragmentation, which is usually treated as the last stage of failure of brittle


materials, constitutes an important role in the completion of material modeling.
When a brittle material is put under stress, the failure process often includes crack
initiation and propagation, and these cracks eventually interact and coalesce to form
fragments. The fragmentation stage of brittle material is important because it is an
efficient energy dissipation mechanism even after the material is considered
“failed” (Johnson and Holmquist 1994). Fragmentation of brittle materials are
thought to be related to strain rate (Hogan et al. 2016), stress state (Hogan et al.
2015), and microstructure (Hogan et al. 2017). Limitations of current models for
fragmentation of brittle materials include: 1. The bulk of the works are focusing on
tensile stress state; 2. Lack of experimental data; and 3. Translating and coupling
fragmentation models into large-scale simulations and advanced micro-mechanical
models. Fragmentation modeling of brittle materials has been extensively reviewed
over the past 15 years. Both (Subhash et al. 2008), focusing on dynamic indentation
fragmentation of ceramics, and (Ramesh et al. 2015), focusing on fragmentation in
planetary and space science, conducted detailed discussions on existing
fragmentation models for brittle materials. In (Subhash et al. 2008), the models
were discussed based on the types and their evolution, whereas in (Ramesh et al.
2015), the emphasis was on examining two important parts of dynamic brittle
fragmentation: the average fragment size, and size distributions. In this sub-section,
three major types of fragmentation models will be reviewed: 1. Geometric statistics-
based models; 2. Energy balance-based models; and 3. Cohesive failure models.
The first two types of models are often called “end-state” models, which describe
the final stage of a fragmentation event (e.g., fragment size and shape distributions).
68 Li et al.

The third type of models is called “process-driven” models, which captures the
earlier states that lead to fragmentation (e.g., crack initiation, propagation, and
coalescence). This sub-section concludes with a brief discussion on the limitations
of these models and potential future directions.

3.3.1 Geometric Statistics-based Fragmentation Models – End-state Models

As the first attempt to understand the fragmentation behavior of brittle


materials, the general objective of geometric statistics-based fragmentation models
was to deduce the fragment size distribution across all length scales (Grady 2010).
These length scales include small-scale deformation mechanisms, failure criterion,
and large-scale loadings. (Åström 2006) conducted a thorough review of the recent
development of statistical models of brittle fragmentation and separated brittle
fragmentation into two groups: 1. Instantaneous fragmentation; and 2. Continuous
fragmentation. In this subsection, the focus will be on instantaneous fragmentation,
which is defined as a process where a dense network of unstable cracks creates
fragments of all sizes within an extremely dynamic stress field.
(Mott and Linfoot 1943) initiated the works on capturing the fragment size
distribution through randomly partitioning geometric shapes with lines on an
infinite area. The theory was developed based on the assumption that energy
requirements in the actual fracture process are negligible and that fracture is
instantaneous. According to the authors, the two-dimensional cumulative fragment
number distribution has the form,

𝑁(𝑎) = 𝑁0 𝑒 −√2𝑁0𝑎 (19)

where a is the fragment area and 𝑁0 is the total fragment numbers. In a later work,
(Grady and Kipp 1985a) modified Mott’s work and treated the one-dimensional
fragmentation problem as a random Poisson point process. On top of the approach
proposed by (Mott and Linfoot 1943), the cumulative fragment number distribution
proposed by (Grady and Kipp 1985a) has the form,

𝑁(𝑎) = 𝑁0 𝑒 −𝑁0 𝑎 (20)

This distribution is called a linear exponential distribution, in contrast to the Mott


distribution, where the argument lies with the exponent being the square root of
fragment area. This representation has also been extended to three-dimensions for
fragment volume estimation (e.g., Gold and Baker 2008). The linear exponential
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 69

distribution has been proven to show improved predictive capability when


compared with Mott distribution. In conjunction with this formulation, different
partitioning approaches at higher dimensions, such as Voronoi tessellation (e.g.,
Du, Faber, and Gunzburger 2005) and Johnson-Mehl construction (e.g., Moller
1992) have been explored.
In (Grady and Kipp 1985b), they also noted that different distributions,
including the Poisson, binomial, log normal, and Weibull fragmentation
distributions, can be obtained using similar geometric approaches. In one of his
later works, (Grady 2008) noted that the fragmentation behavior and fragment
distributions are strongly material dependent, and brittle fragmentation sizes can be
better characterized by a power-law distribution. Multiple approaches employing
both exponential and power-law forms to study fragment size distribution have
been explored in the literature, and details can be referred to in the review by
(Turcotte 1997), (Åström 2006), and (Grady 2009). In general, the power-law
proposed by (Grady 2008) has the form,

𝑁(𝐿) ∝ 𝐿−𝑛 (21)

where L is the fragment size and n is called the fractal dimension. The fractal
dimension has been extensively studied across materials in different areas. (Ramesh
et al. 2015) documented the historical values that have been assigned to n and
concluded that a range of 1.5 to 2.5 is suitable for conventional brittle materials
(e.g., ceramics). However, the determination of fractal dimension highly depends
on the experimental measurements and numerical resolution, and some
characteristic length scales should exist for different materials that bound the
power-law distribution functions.

3.3.2 Energy Balance-based Fragmentation models – End-state Models

While the simple geometric statistics-based models are able to capture the
statistical nature of material fracture and fragmentation, they are not capable of
describing the mechanical nature of fragmentation, which consists of 1. Energy
dissipation during a fragmentation process; and 2. Timescale associated with a
fracture event (Subhash et al. 2008). These limitations have led to the development
of energy balance-based models (e.g., Grady 1982; Glenn and Chudnovskly 1986).
One of the earliest and most widely known models in this class was provided
by (Grady 1982), which considered the local inertial effects in dynamic
fragmentation. This model is governed by an energy balance between the surface
or interface energy created due to material fracture and the local inertia or kinetic
70 Li et al.

energy created due to the fragmentation process. After assuming spherical


fragments of equal size, a simple expression of the fragment diameter or size, d, has
the form (Grady 1982),
2
1 3
(20)2 𝐾𝐼𝐶 (22)
𝑑=[ ]
𝜌𝑐𝜀̇

where 𝜌 is material density, c is the dilatational wave speed, 𝜀̇ is the linear strain
rate, and 𝐾𝐼𝐶 is the material fracture toughness. This model showed good agreement
with the experimentally obtained oil shale and steel data from explosive
fragmentation experiments by the authors. Later, (Glenn and Chudnovskly 1986)
extended the model to account for stored elastic (strain) energy. This was modified
based on the knowledge that the strain energy of a linearly elastic material (i.e.,
ceramics) should be dominant during fragmentation. In the work of (Glenn and
Chudnovskly 1986), the expression of the fragment radius (i.e., similar to the
fragment diameter in Equation (21), a, has the form,

1
𝛼 2 𝜙 (23)
𝑎 = 2 ( ) 𝑠𝑖𝑛ℎ ( )
3 3
3
3 2 (24)
𝜙 = 𝑠𝑖𝑛ℎ−1 𝛽 ( )
𝛼

2𝛽 5 𝜎 ∗ 2
𝛼= + ( ) (25)
𝑅 3 𝜌𝑐𝜀̇

5 𝐾𝐼𝐶 2
𝛽= ( ) (26)
2 𝜌𝑐𝜀̇

where R is the initial radius of the dilating body; 𝜌, c, and 𝜀̇ are the same in Equation
(22) and 𝜎 ∗ is a threshold stress for fragmentation. It is noted by (Glenn and
Chudnovskly 1986) that for a truly brittle material with low fracture toughness and
2𝛽
high fracture-initiation stress, 𝑎 ≈ 𝛼
. One key assumption in this formulation is
that the fragments are stress-free upon their formation so that all the stored elastic
energy is available for fragmentation. However, we emphasize that this is likely the
case only in a tensile stress state, and this assumption was made from an expanding
ring experiment. Whereas in a compression-driven fragmentation event, the
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 71

fragmented pieces will likely remain in contact and interact with each other, from
which a stress-free condition is not likely the case.
In a more recent work by (Grady 2010), a slightly modified version of the
energy expression for average fragment length, 𝜆, has the form,

1
48𝛤 3 (27)
𝜆~ ( 2 )
𝜌𝜀̇

𝐾2
where 𝛤 is the surface fracture energy defined as 𝛤~ 𝜌𝑐𝐼𝐶2. Among Equation (22),
(23), and (27), the characteristic length scales are directly calculated from fracture
properties of a material and a strain rate measurement of the fragmentation event.
It is important to emphasize that for Grady’s energy-based model (Grady 2010), the
experimental data suggested a strong overestimation of the mean fragment size at
low strain rates (~ < 103 s-1). In contrast, this approach seems to provide a good
prediction of the fragment size at high strain rates. By incorporating the strain
energy component, the energy model proposed by (Glenn and Chudnovskly 1986)
gave a much better estimation on the fragment size at quasi-static strain rates. This
indicates the significance of strain energy at low strain rates, where the associated
large timescale in the fragmentation process allows for the accumulation of non-
negligible strain energy in the material. This important consideration of strain
energy has been emphasized and incorporated in (Hogan et al. 2016) towards
compressive loading of advanced ceramics at both quasi-static and dynamic loading
rates. In this work, they modified the models proposed by (Grady and Kipp 1985a)
and (Glenn and Chudnovskly 1986), and model parameters were determined for
silicon carbide, boron carbide, spinel, and basalt. A detailed formulation can be
found in (Hogan et al. 2016). In the modified model by (Hogan et al. 2016), the
stored strain energy in compression was equated to the fracture energy in tension
through an energy-based argument. It has also been noted in their study that the
models are more accurate in the structural-dependent fragmentation region, but not
the microstructure-dependent region (see Figure 6 in Section 2.1.1). From an
experimental perspective, measurements of fragmentation size distributions in
compression were obtained, and this was one of the more comprehensive data set
on fragmentation of ceramics in the literature. This kind of data set would be helpful
in developing compressive-driven fragmentation models. The strain energy
consideration has also been implemented into some further study of rock
fragmentation (Gong et al. 2018; Li et al. 2018) under compression.
72 Li et al.

3.3.3 Cohesive Failure Model – Process-driven Computational Models

Both the geometric statistics-based and energy balance-based fragmentation


models described previously are often categorized as end-state models, which make
use of fragment size and shape data after a fragmentation event happens. While they
can generate fairly accurate predictions, these models cannot capture the
mechanisms associated with the different timescales of processes from fracture to
fragmentation due to their simple formulations. These limitations have led to the
development of computational tools for which the intention is to include both
dynamics and energetics, and account for the residual damage and kinetic energy
of fragments (Subhash et al. 2008). These models are denoted as “process-driven”
models, and the bulk of them are based on one-dimensional wave propagation, and
their solutions are derived using the method of characteristics. In addition, the
majority of these models employ the concept of a cohesive failure model in which
failure is captured through an explicit relation between the cohesive traction acting
across some microstructural heterogeneity (e.g., microcrack) and the resulting
displacement differences. Naturally, these models are an extension of the previous
energy balance-based models, where the whole event from fracture initiation (i.e.,
crack nucleation) to fragmentation can be captured in a physics-based framework.
One earlier work was proposed by (Miller, Freund, and Needleman 1999) using
cohesive elements in a one-dimension finite element scheme to investigate dynamic
brittle fragmentation. They incorporate time history of the fracture and
fragmentation process, and the number, distribution, and strength of flaws in the
material. The numerical simulation predicted fragment size an order of magnitude
smaller than was predicted by the energy balanced-based models (e.g., Grady 1982;
Glenn and Chudnovskly 1986). They argued that this large difference was the result
of the missing time dependency for crack nucleation and propagation over a finite
time interval, as the energy balance-based models only accounted for the onset.
Shown in Figure 49 is an example of the change in energies during the
fragmentation process in (Miller, Freund, and Needleman 1999) demonstrating the
time evolution of various energies.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 73

Figure 49: Time histories of energy contributions: (a) Up to the time of fragmentation, (b)
During the entire computation time. WAP is the work done by the imposed loading, WEL is the
word done by the elastic strain energy, KE is the kinetic energy, and 𝛷 is the cohesive energy.
These images are taken from (Miller, Freund, and Needleman 1999).

A later work by (Drugan 2001) was dedicated to accounting for wave


interactions for one-dimensional fragmentation of a bar. He studied two cases: one
with unflawed material and the other with initial flaw spacing. The critical outcome
of his study was that while the aforementioned simple energy balance-based models
were found to agree with his simulations for extremely high strain rates (in excess
of 5 x 107 s-1 for dense alumina), they consistently predict much larger fragment
sizes at lower strain rate regimes. Shown in Figure 50 is a comparison between the
simulations by (Drugan 2001) and the energy balanced-based models. Some
important points can be made from this figure: 1. This model indicates that the
fragment size is strain-rate independent in the quasi-static regime; 2. The energy
balanced-based model clearly overestimates the fragment size at low strain rates; 3.
This model agrees well with the model presented by (Miller, Freund, and
Needleman 1999), who used a different approach to solve the problem.
74 Li et al.

Figure 50: Comparison of several models’ predictions of fragment size versus applied
strain rate for dense alumina. The simulation results of Drugan model are shown as black dots.
The closed-form solution of Drugan model in the quasi-static strain rate regime is shown as the
horizontal line segment. This image is taken from (Drugan 2001).

The extension of Drugan’s framework was presented by (Shenoy and Kim


2003). Their modifications included: 1. Incorporating a dynamic mean-field theory;
2. Accounting for the statistical nature of flaw distributions; and 3. Accounting for
stress wave interactions between flaws. They showed that their model gave an
accurate description of the damage process and predicted fragment sizes that are in
close quantitative agreement with the numerical calculations.
More recent work based on similar concepts, but incorporating elastic wave
propagation and interactions, crack nucleation and growth, as well as defect
distributions in a fragmentation simulation were presented by (Zhou, Molinari, and
Ramesh 2005, 2006b, 2006a) (ZMR) for a much wider range of strain rates and
materials. In (Zhou, Molinari, and Ramesh 2005), a one-dimensional bar problem
was solved, and in (Zhou, Molinari, and Ramesh 2006a) the problem was extended
for a two-dimensional circular ring that was dynamically expanded. The models
incorporated a dynamic crack initiation criterion and a cohesive crack growth
model that were used to describe material fracture and failure. Shown in Figure 51
is an illustration of the cohesive crack growth model and a comparison between the
ZMR model and the energy balanced-based models. The ZMR model agrees well
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 75

with the Drugan model at intermediate strain rates (~ 5 x 104 s-1 to ~ 2 x 105 s-1) but
predicts more than twice larger fragment size than the Drugan model at quasi-static
strain rates and three to five times smaller fragment size at extremely high strain
rates (~ 107 s-1). They also demonstrated that the initial flaw distribution had a
significant effect on fragment size distributions in the strain rates investigated.
Similar observations were made by (Levy and Molinari 2010), and they
incorporated a communication factor into their scaling function for the average
fragment size across different flaw distributions.

Figure 51: (a) An irreversible softening cohesive model describes the crack
opening/closing process; (b) A comparison of Average fragment size vs. Strain rate
(homogeneous bar without defects) between different models. These images are taken from
(Zhou, Molinari, and Ramesh 2006a).

In the same year, (Zhou, Molinari, and Ramesh 2006b) studied the effect of
material properties on the fragmentation of brittle materials and concluded with a
strain rate-dependent normalized scheme for all the materials studied. A rather
surprising but meaningful outcome was generated in this study. They found that all
of the material behaviors were shown to collapse to a single rate-dependent curve
for the rate-dependent average fragment size. A characteristic time was defined in
(Zhou, Molinari, and Ramesh 2006b),

𝐸𝐺𝑐
𝑡0 = (28)
𝑐𝜎𝑐2

The characteristic length was defined as,

𝐿0 = 𝑐𝑡0 (29)

The characteristic strain rate was defined as,


76 Li et al.

𝑐𝜎𝑐3
𝜀̇0 = (30)
𝐸 2 𝐺𝑐

where 𝐺𝑐 is the fracture energy, 𝜎𝑐 is a critical strength of the material (i.e., similar
to that of 𝜎 ∗ in Equation (7)), E is the Young’s modulus, and c is the dilatational
wave speed. Finally, a general expression of the average fragment size, s, is related
to,
𝑠 = 𝑓(𝐸, 𝑐, 𝜎𝑐 , 𝐺𝑐 , 𝜀̇) (31)

In non-dimensional form, the average fragment size, 𝑠̅ , has the form,

𝜁
𝜎𝑐 𝛾+2−3𝜁 𝐸 2 𝐺𝑐 𝜀̇
𝑠̅ = 𝑓̅ [( ) ( ) ] = 𝑓̅(𝜎̅𝑐 , 𝜀̇)̅ (32)
𝐸 𝑐𝜎𝑐3

where 𝛾 and 𝜁 are the assumed power of 𝜎𝑐 and 𝜀̇ in non-dimensional analysis,


𝜎
respectively; 𝜎̅𝑐 = 𝑐 is the non-dimensional critical stress. This was further
𝐸
simplified by assuming the critical strength has a negligible effect (at least in the
data set that (Zhou, Molinari, and Ramesh 2006b) used), and the average fragment
size was only strain rate-dependent. Therefore, combining the power law of 2/3
proposed by (Grady 1982) (see Equation ((22)), the function could have the form,

𝑝1
𝑠̅ = 2 (33)
1 + 𝑎𝜀̇ 3

where 𝑝1 and 𝑎 are fitted coefficients. Some more accurate fits were also provided
in the study, which will be omitted here. Figure 52 shows the compiled curve
generated by assuming different fracture energies, Young’s moduli, and material
wave speed, which are the material properties of different brittle materials. Note
that these simulation results are the outcome of using the ZMR model. More
sophisticated simulations (e.g., Levy and Molinari 2010; Bažant and Caner 2014)
built on these works and provided a further understanding of the scaling effect of
fragmentation with respect to strain rate. It is also worth mentioning that (Levy and
Molinari 2010) found that the number of large defects and the rate at which cracks
can be initiated have a profound effect on the stress release waves, and
consequently, the average fragment size.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 77

Figure 52: Calculated non-dimensional fragment size vs. non-dimensional strain-rate for
all of the material properties examined in (Zhou, Molinari, and Ramesh 2006b). The
predictions of other theories are included for comparison (i.e., Fit 1 and 2). Details of the fitted
equations can refer to (Zhou, Molinari, and Ramesh 2006b). This image is taken and modified
from (Zhou, Molinari, and Ramesh 2006b).

Besides modeling the “macroscopic” fragmentation events, some rather


distinctive insights have been provided from a different group of researchers that
studied brittle fragmentation from a grain-level perspective (e.g., Espinosa,
Zavattieri, and Dwivedi 1998; Zavattieri, Raghuram, and Espinosa 2001; Zavattieri
and Espinosa 2001; Maiti, Rangaswamy, and Geubelle 2005; Kraft et al. 2008).
These explicit models adapted intergranular cohesive failure models and
intragranular damage models to probe the effect of grains (i.e., crack propagation,
grain-grain interaction) on fragment size. (Maiti, Rangaswamy, and Geubelle 2005)
showed the influence of average grain size on the final damage and the average
fragment size of the material under high-rate loading. (Kraft et al. 2008) modeled
intergranular cracking using a distribution of cohesive interfaces and reported the
effect of confinement, grain boundary friction, and spatial flaw distribution on
strength. A trend of increasing failure strength with decreasing flaw density was
demonstrated. However, it should be noted that a general consideration in using the
cohesive model is the proper choice of model parameters as the cohesive failure
law is phenomenological in nature, given the fact that the cohesive zone of brittle
materials is typically much smaller than for ductile materials. These parameters
could be quantitatively measured, for example, by coupling both high-resolution x-
78 Li et al.

ray tomography and digital image correlation at a fracture zone in the material
(Viggiani et al. 2007). In addition, these studies pioneered the exploration of
material microstructure effects on the fragmentation process. The role of
microstructure has been proven to be important in brittle fragmentation (see Section
2 of this chapter), especially under compressive stress states.
More recently, efforts by (Cereceda, Graham-Brady, and Daphalapurkar 2017;
Cereceda et al. 2018) on modeling dynamic fragmentation of heterogeneous brittle
materials showed promising advancements in the extension of one-dimensional bar
simulation to two-dimension or three-dimensional finite element analysis (FEA)-
based simulations. The model of (Cereceda, Graham-Brady, and Daphalapurkar
2017) adapted the ZMR one-dimensional model and integrated a localized strain
rate formulation to better capture the fragment size distribution of three modeled
materials: 1. glass; 2. concrete; and 3. masonry. Shown in Figure 53 is a comparison
of the model prediction with and without the heterogeneity of strain rates. Note that
this model incorporated spatially varying strength to represent the heterogeneity in
material microstructures. The model did not capture the probability density function
of the experimental data as it was, but significant improvements were made when
a heterogeneous strain rate formulation was defined. The detailed formulation can
be referred to (Cereceda, Graham-Brady, and Daphalapurkar 2017), which is
omitted here.

Figure 53: (a) Comparison of normalized fragment-size probability distribution function


from 1D simulations with 3D shock tube experiments in masonry, concrete and glass at a strain
rate of 104 s-1 with a standard deviation of strength of 4%; (b) Comparison of fragment-size pdf
from 1D simulations with 3D shock tube experiments in masonry when using a heterogeneous
strain rate formulation based on weight functions. The fragment-size threshold is sth = 0.6 mm.
These images are taken from (Cereceda, Graham-Brady, and Daphalapurkar 2017).

Efforts were devoted in one of their later studies (Cereceda et al. 2018) to model
the dynamic fragmentation of a multi-phase material, where masonry (i.e., the
matrix (brick), inclusions (mortar), and interfaces) was used as modeled material in
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 79

a micro-mechanical approach. Similar to their previous study (Cereceda, Graham-


Brady, and Daphalapurkar 2017), a critical strength distribution and a cohesive
zone model were used to model heterogeneity and progressive damage,
respectively. The influence of inclusion volume fraction and size on the average
fragment-size and fragment-size distributions were found for strain rates between
102 s-1 and 105 s-1. Their results suggest a critical inclusion size exists, which
depends on strain rate and material properties, below which no additional fragments
of the inclusion-material are created.
As mentioned at the beginning of this sub-section, one limitation of the
aforementioned fragmentation models is that they all assume a tensile stress state
to start with. This becomes unrealistic when one considers most of the structural
applications for advanced ceramics (e.g., impact), in which primarily compressive
stress states are generated. The fragment-to-fragment contact is dominant in the
case of compression-driven dynamic fragmentation, especially under confinement
(Subhash et al. 2008). This issue is further complicated when more complex
fracture initiation criterion, fracture patterns (e.g., wing crack), crack-crack
interaction and coalescence, and inelasticity are taken into consideration.
Furthermore, the effect of microstructure always plays an important role in
fragmentation (Hogan et al. 2015; Hogan et al. 2016). Even though numerous
micro-mechanical models have been developed to capture how these features
emerge, they are not able to translate these fracture events directly into
fragmentation. This decoupled approach (i.e., fragment sizes are estimated by
stopping the simulation at some ad hoc time, estimating the strain rate distribution,
and then using post-processing calculations to extract fragment distributions) has
the major disadvantage of providing solutions that depend on the time chosen to
estimate the onset of fragmentation (Ramesh et al. 2015). Thus, future development
of micro-mechanical models towards brittle materials should aim at including crack
nucleation and coupling of fragmentation models to the fracture models. Further,
the incorporation of further granular flow after fragmentation are needed, as it has
been noted that particle granular flow persists significant effects on post-
fragmentation events such as projectile erosion (Cil, Hurley, and Graham-Brady
2019).
Another note before concluding this sub-section is that there are no specific
fragmentation models which have been developed for boron carbide. The
fragmentation behavior of boron carbide is expected to fall into the general
prediction for advanced ceramics. In fact, many experimental and modeling works
have used boron carbide data for model fitting and validation (e.g., Moynihan,
LaSalvia, and Burkins 2002; Grady 2010; Tonge and Ramesh 2016b; Cil, Hurley,
and Graham-Brady 2019), and reasonable agreements have been obtained.
80 Li et al.

However, the authors expect that if the unique fracture and fragmentation
mechanisms of boron carbide (i.e., amorphization-induced fracture, transgranular
cracking) can be specifically incorporated, a model can generate more accurate and
insightful results for boron carbide. In the meantime, more experimental works
should be done on boron carbide to generate data at different stress states and strain
rates so that an enhanced understanding on the activation of these mechanisms, and
how they relate to fragmentation behavior can be gained.

4. FUTURE DIRECTIONS

In this final section, we identify future directions for mechanical testing and
modeling of the fracture and fragmentation of boron carbide. The pressing issues
related to the current testing methods and modeling efforts will be re-emphasized,
and recent techniques and modeling concepts which could potentially enhance our
understanding of the fracture and fragmentation behavior of boron carbide will be
mentioned. In general, the overall pursuits of future research will be centered
around capturing the transitions from intact material to damaged material, then from
damaged material to fragmented material. Altogether, this provides a roadmap
towards obtaining a more complete picture of the boron carbide fracture and
fragmentation mechanisms for use in the design of improved materials and
structures in industrial and defense applications.
From an experimental mechanics perspective, even though ultra-high-speed
imaging has played an important role in the real-time visualization of material
fracture and fragmentation, higher resolution at faster recording rates are still
desired to resolve dynamic fracture and fragmentation events that happen within a
few microseconds. This will be especially useful in extremely high strain rate
impact loading (e.g., ballistic impact, laser shock). Another limitation of high-speed
imaging is that the specimen is usually out of focus when fragmentation occurs due
to the limited depth-of-field. Thus, implementation of real-time diagnostic
techniques, for example, the photonic Doppler velocimetry (PDV) system (e.g.,
Tonge and Schuster 2017; Mallick et al. 2019), could further enhance our
understanding of the time-resolved fragmentation behavior. To probe the
transitional behavior from intact material to damaged material, in-situ x-ray
computed tomography (e.g., Buffiere et al. 2010), in-situ phase-contrast imaging
(e.g., Leong et al. 2018), and in-situ acoustic measurements of fracture (e.g.,
Eberhardt et al. 1997) are favorable if they can be implemented into dynamic and
impact testing. In addition, a recent investigation has focused on using in-situ x-
ray diffraction on shock-compressed silicon carbide to determine the onset of
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 81

inelastic deformation and phase transition (Tracy et al. 2019). The combination of
these in-situ techniques could bridge the gap between transitional phases (i.e.,
intact-damage and damage-fragmentation) which cannot be accessed in a
conventional test, and the results obtained can be used to better understand
postmortem measurements of fracture and fragmentation (e.g., Hogan, Farbaniec,
Sano, et al. 2016; Hogan et al. 2017).
Coupled with in-situ diagnostics is the use of improved post-imaging analysis
or post-mortem characterization tools. Digital image correlation has been starting
to be employed to look at these transitional behaviors, where crack initiation and
propagation can be qualitative and quantitative measured and correlated with
localized strain evolution (Cidade et al. 2019). The advantages of digital image
correlation include: 1. Resolving two-dimensional or three-dimensional spatial
resolution of strains; 2. Capable of capturing localized features (e.g., crack initiation
and propagation) and correlate them with strain-time history profile; 3. Generating
spatial-temporal strain data (e.g., strain-time history profile at localized features);
and 4. Low cost, non-destructive method compared to conventional strain gauges.
Overall, we emphasize coupling advanced diagnostic techniques onto the existing
experimental setups focusing on the transitional behaviors among the three stages
of boron carbide failure, from which the data generated will be useful in model
implementation and validation. Some examples include in-situ SEM (e.g., Song et
al. 2008; Romeis et al. 2012; Palomares-García, Pérez-Prado, and Molina-
Aldareguia 2017) and in-situ TEM (e.g., De Hosson 2009; C. Q. Chen, Pei, and De
Hosson 2010; Kiener and Minor 2011; Ye et al. 2011) coupled with micropillar
compression, nanoindentation, tensile testing, and bending. In addition, the recent
development in aberration-corrected high-resolution TEM provides even higher
resolution, for which nanoscale features can be more thorough studied. At the same
time, it is also crucial to generate more data for boron carbide under impact loading
conditions, where the lack of experimental works is significantly limiting the pace
of model development.
From a modeling perspective, current computing capacity limits the study of
multi-length scale material response at a realistic physical dimension. Thus, new
modeling concepts, including the incorporation of mathematical tools for scale
transitioning, homogenization, and handling numerical and material instabilities
could be extremely helpful. The specific goal with modeling will be to generate a
robust multi-scale micro-mechanical model that links material microstructure and
the lowest scale deformation mechanisms (e.g., dislocation and twinning) to large
scale fracture (e.g., wing crack), then to macroscopic material behaviors (e.g.,
fragmentation/granular flow). This includes improvements in, for example,
meshing strategies for multi-scale fracture (e.g., microscopic cracking to structural
82 Li et al.

fracture), nanoscale inelastic deformation to microscale fracture (e.g.,


amorphization to cracking), and interactions between these different mechanisms.
We emphasize, again, the need for both implicit and explicit modeling of the
transitional behaviors of failure (i.e., intact-fracture, then fracture to fragmentation),
for which even some of the most recent micro-mechanical models have not
incorporated (e.g., Clayton and Knap 2018; Zeng, Tonge, and Ramesh 2019). In
addition, the lack of communication between micro-mechanical fracture and
fragmentation models limits the accuracy in predicting material post-damage
behavior. Furthermore, expanding the current micro-mechanical fracture models by
incorporating deformation-induced flaws (e.g., activation of slip systems) over
processing-induced flaws (e.g., grain boundary phases) may improve modeling
accuracy and application to materials design. In the meantime, broadening current
fragmentation models by incorporating history-dependent load-induced damage
over the prescribed flaw distribution for micro-crack nucleation and coalescence
remains a major challenge in compression-driven fragmentation. Furthermore,
some new modeling approaches, such as molecular dynamics simulations, are not
heavily invested in studying the fracture and fragmentation of boron carbide yet
(e.g., Taylor, McCauley, and Wright 2012). The investigation at even smaller
length scales (e.g., atomic) may provide further insights on the formation of
amorphization and how it relates to fracture. Beyond current mechanics approaches,
the use of machine learning for material design and solving probabilistic problems
(e.g., resolving fragmentation distributions using a set of microstructures) has been
brought up recently (e.g., Liu et al. 2017; Ward et al. 2018; Ying Zhang and Ling
2018) and may be worth pursuing in the future for studying fracture and
fragmentation of boron carbide.
Finally, it is worth re-emphasizing that inefficient and inaccurate validation
limits model development. Through this review, the authors would like to raise
some questions towards future experimental and model validation: 1. How do we
compare complex fracture and fragmentation patterns with simulation results? 2.
How can we improve experimental validation techniques to more accurately match
models (e.g., quantitatively measure the parameters in a cohesive zone model)? 3.
How to perform complex history-dependent experiments in order to better validate
model outputs? 4. How should we model and quantify the variability in the data
generated during experiments (e.g., dynamic compressive strength)? Answering
these questions will significantly improve the efficiency in communication between
experiments and models. One complimentary solution for the above questions is to
have an openly available and comprehensive experimental, theoretical, and
computational data sets for modeling fracture and fragmentation of boron carbide.
With such an open data set, we hope that the collaboration between experimental
mechanics and modeling can be more fruitful.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 83

5. CONCLUSION
We have presented an overview of the experimental and theoretical studies
focused on the fracture and fragmentation behavior of boron carbide. In order to
characterize the dynamic response of advanced ceramics over a wide range of
loading conditions, a variety of experimental techniques have been developed.
Consequently, the fracture and fragmentation behavior of boron carbide has been
probed across a range of stress states and strain rates, from the controlled uniaxial
compressive stress states of split-Hopkinson pressure bar tests to the complex stress
states and high strain rates in ballistic experiments. The quantitative data and
insights generated from experimental studies inform our understanding of the
physical mechanisms behind the behavior of brittle materials and have facilitated
the development of computational models for brittle fracture and fragmentation.
Such models play a key role in furthering our understanding of fracture and
fragmentation processes, ultimately guiding the development of improved boron
carbide materials and structures that incorporate them.

REFERENCES

Alekseevskii, VP. 1966. “Penetration of a Rod into a Target at High Velocity.”


Combustion, Explosion, and Shock Waves 2 (2): 63–66.
Anderson, Charles E., Matthew S. Burkins, James D. Walker, and William A. Gooch.
2005. “Time-Resolved Penetration of B 4C Tiles by the APM2 Bullet.” CMES -
Computer Modeling in Engineering and Sciences 8 (2): 91–104.
Anderson, Charles E., and Suzanne A. Royal-Timmons. 1997. “Ballistic Performance
of Confined 99.5%-Al203 Ceramic Tiles.” International Journal of Impact
Engineering 19 (8): 703–13. https://doi.org/10.1016/s0734-743x(97)00006-7.
Anderson, Charles E., and James D Walker. 2005. “An Analytical Model for Dwell and
Interface Defeat.” International Journal of Impact Engineering 31 (9): 1119–32.
https://doi.org/10.1016/j.ijimpeng.2004.07.013.
Anderson, Charles E, and Bruce L Morris. 1992. “The Ballistic Performance of
Confined Al2o3 Ceramic Tiles.” International Journal of Impact Engineering 12
(2): 167–87.
Anderson Jr, Charles E. 2009. “Compression Testing and Response of SiC-N Ceramics:
Intact, Damaged and Powder.” In Advances in Ceramic Armor: A Collection of
Papers Presented at the 29th International Conference on Advanced Ceramics
and Composites, Jan 23-28, 2005, Cocoa Beach, FL, 109. John Wiley & Sons,
Inc.
Anderson Jr, Charles E, and James D Walker. 1991. “An Examination of Long-Rod
Penetration.” International Journal of Impact Engineering 11 (4): 481–501.
Anderson Jr, Charles E, James D Walker, and Jim Lankford. 1995. “Investigations of
the Ballistic Response of Brittle Materials.” San Antonio, TX.
Arrowood, Roy, and James Lankford. 1987. “Compressive Fracture Processes in an
84 Li et al.

Alumina-Glass Composite.” Journal of Materials Science 22 (10): 3737–44.


https://doi.org/10.1007/BF01161487.
ARSLAN, Kemal, and Recep GÜNEŞ. 2017. “Ballistic Impact Simulation of
Ceramic/Metal Armor Structures.” Uluslararası Muhendislik Arastirma ve
Gelistirme Dergisi, 12–18. https://doi.org/10.29137/umagd.371100.
Ashby, M. F., and C. G. Sammis. 1990. “The Damage Mechanics of Brittle Solids in
Compression.” Pure and Applied Geophysics 133 (3): 489–521.
Åström, J A. 2006. “Statistical Models of Brittle Fragmentation.” Advances in Physics
55 (3–4): 247–78. https://doi.org/10.1080/00018730600731907.
Ayyagari, R. S., N. P. Daphalapurkar, and K. T. Ramesh. 2018. “The Effective
Compliance of Spatially Evolving Planar Wing-Cracks.” Journal of the
Mechanics and Physics of Solids 111: 503–29.
https://doi.org/10.1016/j.jmps.2017.11.016.
Bavdekar, Salil, Gregory Parsard, Ghatu Subhash, and Sikhanda Satapathy. 2017. “An
Improved Dynamic Expanding Cavity Model for High-Pressure and High-Strain
Rate Response of Ceramics.” International Journal of Solids and Structures 125:
77–88. https://doi.org/10.1016/j.ijsolstr.2017.07.014.
Bavdekar, Salil, Ghatu Subhash, and Sikhanda Satapathy. 2019. “A Unified Model for
Dwell and Penetration during Long Rod Impact on Thick Ceramic Targets.”
International Journal of Impact Engineering 131 (May): 304–16.
https://doi.org/10.1016/j.ijimpeng.2019.05.014.
Bažant, Zdeněk P., and Ferhun C Caner. 2014. “Impact Comminution of Solids Due to
Local Kinetic Energy of High Shear Strain Rate: I. Continuum Theory and
Turbulence Analogy.” Journal of the Mechanics and Physics of Solids 64 (1):
223–35. https://doi.org/10.1016/j.jmps.2013.11.008.
Betranhandy, Emmanuel, Nathalie Vast, and Jelena Sjakste. 2012. “Ab Initio Study of
Defective Chains in Icosahedral Boron Carbide B 4C.” In Solid State Sciences,
14:1683–87. Elsevier Masson SAS.
https://doi.org/10.1016/j.solidstatesciences.2012.07.002.
Bhagavathula, Kapil Bharadwaj, Austin Azar, Simon Ouellet, Sikhanda Satapathy,
Christopher R. Dennison, and James David Hogan. 2018. “High Rate
Compressive Behaviour of a Dilatant Polymeric Foam.” Journal of Dynamic
Behavior of Materials 4 (4): 573–85. https://doi.org/10.1007/s40870-018-0176-
0.
Birkhoff, Garrett, Duncan P. MacDougall, Emerson M Pugh, and Geoffrey Taylor.
1948. “Explosives with Lined Cavities.” Journal of Applied Physics 19 (6): 563–
82. https://doi.org/10.1063/1.1698173.
Born, M. 1942. “On the Stability of Crystal Lattices.” Proc Camb Phil Soc, no.
September 1939: 160–72.
Bouchaud, Elisabeth, Dominique Jeulin, Claude Prioul, and Stephane Roux. 2012.
Physical Aspects of Fracture. Springer Science and Business Media.
Bourne, N. K. 2002. “Shock-Induced Brittle Failure of Boron Carbide.” Proceedings of
the Royal Society A: Mathematical, Physical and Engineering Sciences 458
(2024): 1999–2006. https://doi.org/10.1098/rspa.2002.0968.
Brace, WF, and EG Bombolakis. 1963. “A Note on Brittle Crack Growth in
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 85

Compression.” Journal of Geophysical Research 68 (12): 3709–13.


Budiansky, Bernard, and Richard J. O’connell. 1976. “Elastic Moduli of a Cracked
Solid.” International Journal of Solids and Structures 12 (2): 81–97.
https://doi.org/10.1016/0020-7683(76)90044-5.
Buffiere, J. Y., E. Maire, J. Adrien, J. P. Masse, and E. Boller. 2010. “In Situ
Experiments with X Ray Tomography: An Attractive Tool for Experimental
Mechanics.” In Proceedings of the Society for Experimental Mechanics, Inc.,
67:289–305. https://doi.org/10.1007/s11340-010-9333-7.
Cereceda, David, Lori Graham-Brady, and Nitin Daphalapurkar. 2017. “Modeling
Dynamic Fragmentation of Heterogeneous Brittle Materials.” International
Journal of Impact Engineering 99: 85–101.
https://doi.org/10.1016/j.ijimpeng.2016.09.012.
Cereceda, David, Dmitriy Kats, Nitin Daphalapurkar, and Lori Graham-Brady. 2018.
“A Micro-Mechanical Modeling Approach for Dynamic Fragmentation in Brittle
Multi-Phase Materials.” International Journal of Solids and Structures 134: 116–
29. https://doi.org/10.1016/j.ijsolstr.2017.10.026.
Chen, C Q, Y T Pei, and Jeff Th M De Hosson. 2010. “Effects of Size on the Mechanical
Response of Metallic Glasses Investigated through in Situ TEM Bending and
Compression Experiments.” Acta Materialia 58 (1): 189–200.
https://doi.org/10.1016/j.actamat.2009.08.070.
Chen, R., W. Yao, F. Lu, and K. Xia. 2018. “Evaluation of the Stress Equilibrium
Condition in Axially Constrained Triaxial SHPB Tests.” Experimental
Mechanics 58 (3): 527–31. https://doi.org/10.1007/s11340-017-0344-5.
Chen, W., B. Song, D. J. Frew, and M. J. Forrestal. 2003. “Dynamic Small Strain
Measurements of a Metal Specimen with a Split Hopkinson Pressure Bar.”
Experimental Mechanics 43 (1): 20–23. https://doi.org/10.1007/BF02410479.
Chen, W., and Bo Song. 2010. Split Hopkinson (Kolsky) Bar: Design, Testing and
Applications. Springer Science & Business Media.
Chen, Weinong, and G. Ravichandran. 2000. “Failure Mode Transition in Ceramics
under Dynamic Multiaxial Compression.” International Journal of Fracture 101
(1/2): 141–59. https://doi.org/10.1023/A:1007672422700.
Chen, Weinong, and G Ravichandran. 1997. “Dynamic Failure of a Glass Ceramic.”
Journal of Mechanics and Physics of Solids 45 (8): 1303–28.
Chocron, Sidney, Charles E. Anderson, Kathryn A. Dannemann, Arthur E. Nicholls,
and Nikki L. King. 2012a. “Intact and Predamaged Boron Carbide Strength under
Moderate Confinement Pressures.” Journal of the American Ceramic Society 95
(1): 350–57. https://doi.org/10.1111/j.1551-2916.2011.04931.x.
Chocron, Sidney, Charles E. Anderson, Kathryn A Dannemann, Arthur E Nicholls, and
Nikki L King. 2012b. “Intact and Predamaged Boron Carbide Strength under
Moderate Confinement Pressures.” Journal of the American Ceramic Society 95
(1): 350–57. https://doi.org/10.1111/j.1551-2916.2011.04931.x.
Cicchetti, Nicole A, Bazle Z Haque, Shridhar Yarlagadda, and John W Gillespie Jr.
2014. “Modeling and Simulation of Ceramic Arrays to Improve Ballistic
Performance.” Newark.
Cidade, Rafael A, Daniel S.V. Castro, Enrique M Castrodeza, Peter Kuhn, Giuseppe
86 Li et al.

Catalanotti, Jose Xavier, and Pedro P Camanho. 2019. “Determination of Mode I


Dynamic Fracture Toughness of IM7-8552 Composites by Digital Image
Correlation and Machine Learning.” Composite Structures 210 (June 2018): 707–
14. https://doi.org/10.1016/j.compstruct.2018.11.089.
Cil, Mehmet B, Ryan C Hurley, and Lori Graham-Brady. 2019. “A Rate-Dependent
Constitutive Model for Brittle Granular Materials Based on Breakage
Mechanics.” Journal of the American Ceramic Society, no. October 2018: 5524–
34. https://doi.org/10.1111/jace.16376.
Clayton, J. D. 2012. “Towards a Nonlinear Elastic Representation of Finite
Compression and Instability of Boron Carbide Ceramic.” Philosophical
Magazine 92 (23): 2860–93. https://doi.org/10.1080/14786435.2012.682171.
———. 2013. “Mesoscale Modeling of Dynamic Compression of Boron Carbide
Polycrystals.” Mechanics Research Communications 49: 57–64.
https://doi.org/10.1016/j.mechrescom.2013.02.005.
———. 2014. “Phase Field Theory and Analysis of Pressure-Shear Induced
Amorphization and Failure in Boron Carbide Ceramic.” AIMS Materials Science
1 (3): 143–58. https://doi.org/10.3934/matersci.2014.3.143.
———. 2015. “Penetration Resistance of Armor Ceramics: Dimensional Analysis and
Property Correlations.” International Journal of Impact Engineering 85: 124–31.
https://doi.org/10.1016/j.ijimpeng.2015.06.025.
———. 2016. “Dimensional Analysis and Extended Hydrodynamic Theory Applied to
Long-Rod Penetration of Ceramics.” Defence Technology 12 (4): 334–42.
https://doi.org/10.1016/j.dt.2016.02.004.
Clayton, J. D., and J. Knap. 2018. Continuum Modeling of Twinning, Amorphization,
and Fracture: Theory and Numerical Simulations. Continuum Mechanics and
Thermodynamics. Vol. 30. Springer Berlin Heidelberg.
https://doi.org/10.1007/s00161-017-0604-8.
Cottrell, M G, J. Yu, Z. J. Wei, and D. R.J. Owen. 2003. “The Numerical Modelling of
Ceramics Subject to Impact Using Adaptive Discrete Element Techniques.”
Engineering Computations (Swansea, Wales) 20 (1–2): 82–106.
https://doi.org/10.1108/02644400310458856.
Cronin, Duane S, Khahn Bui, Christian Kaufmann, Grant Mcintosh, and Todd Berstad.
2004. “Implementation and Validation of the Johnson-Holmquist Ceramic
Material Model in LS-Dyna.” 4th European LS-DYNA Users Conference, 47–60.
Crouch, Ian G., Gareth Appleby-Thomas, and Paul J. Hazell. 2015. “A Study of the
Penetration Behaviour of Mild-Steel-Cored Ammunition against Boron Carbide
Ceramic Armours.” International Journal of Impact Engineering 80: 203–11.
https://doi.org/10.1016/j.ijimpeng.2015.03.002.
Cui, Feng-Dan, Tian Ma, and Wei-Ping Li. 2017. “Damage Characteristics of SiC and
B4C Ballistic Insert Plates Subjected to Multi-Hit.” Journal of Inorganic
Materials Beijing 32 (9): 967–72.
Curran, D. R., L. Seaman, T. Cooper, and D. A. Shockey. 1993. “Micromechanical
Model for Comminution and Granular Flow of Brittle Material under High Strain
Rate Application to Penetration of Ceramic Targets.” International Journal of
Impact Engineering 13 (1): 53–83. https://doi.org/10.1016/0734-
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 87

743X(93)90108-J.
Dandekar, Dattatraya P. 2001. “Shock Response of Boron Carbide.” Aberdeen Proving
Ground MD.
Deng, H., and S. Nemat-Nasser. 1992. “Dynamic Damage Evolution in Brittle Solids.”
Mechanics of Materials 14 (2): 83–103. https://doi.org/10.1016/0167-
6636(92)90008-2.
Deng, H, and S. Nemat-Nasser. 1994. “Strain-Rate Effect on Brittle Failure in
Compression.” Acta Metallurgica et Materialia 42 (3): 1013–24.
Deshpande, V. S., and A. G. Evans. 2008. “Inelastic Deformation and Energy
Dissipation in Ceramics: A Mechanism-Based Constitutive Model.” Journal of
the Mechanics and Physics of Solids 56 (10): 3077–3100.
https://doi.org/10.1016/j.jmps.2008.05.002.
DeVries, Matthew, John Pittari, Ghatu Subhash, Kendall Mills, Chris Haines, James Q.
Zheng, and F. Zok. 2016a. “Rate-Dependent Mechanical Behavior and
Amorphization of Ultrafine-Grained Boron Carbide.” Journal of the American
Ceramic Society 99 (10): 3398–3405. https://doi.org/10.1111/jace.14324.
———. 2016b. “Rate-Dependent Mechanical Behavior and Amorphization of
Ultrafine-Grained Boron Carbide.” Journal of the American Ceramic Society 99
(10): 3398–3405. https://doi.org/10.1111/jace.14324.
Dolan, D.H. 2006. “Foundations of VISAR Analysis.” Albuquerque, NM.
https://doi.org/10.2514/6.1985-1421.
Domnich, Vladislav, Yury Gogotsi, Michael Trenary, and Takaho Tanaka. 2002.
“Nanoindentation and Raman Spectroscopy Studies of Boron Carbide Single
Crystals.” Applied Physics Letters 81 (20): 3783–85.
https://doi.org/10.1063/1.1521580.
Domnich, Vladislav, Sara Reynaud, Richard A. Haber, and Manish Chhowalla. 2011.
“Boron Carbide: Structure, Properties, and Stability under Stress.” Journal of the
American Ceramic Society 94 (11): 3605–28. https://doi.org/10.1111/j.1551-
2916.2011.04865.x.
Donath, Fred A. 1961. “Rocks, Experimental Study of Shear Failure in Anisotropic.”
Geological Society of America Bulletin 72: 985–89.
Drugan, W J. 2001. “Dynamic Fragmentation of Brittle Materials : Analytical
Mechanics-Based Models.” Journal of the Mechanics and Physics of Solids 49:
1181–1208.
Du, Qiang, Vance Faber, and Max Gunzburger. 2005. “Centroidal Voronoi
Tessellations: Applications and Algorithms.” SIAM Review 41 (4): 637–76.
https://doi.org/10.1137/s0036144599352836.
Dyachkov, S. A., A. N. Parshikov, M. S. Egorova, S. Yu Grigoryev, V. V. Zhakhovsky,
and S. A. Medin. 2018. “Explicit Failure Model for Boron Carbide Ceramics
under Shock Loading.” Journal of Applied Physics 124 (8).
https://doi.org/10.1063/1.5043418.
Eberhardt, E, D Stead, B Stimpson, R S Read, Geological Engineering, and Whiteshell
Laboratories. 1997. “Changes in Acoustic Event Properties with Progressie
Fracture Damage.” International Journal of Rock Mechanics and Mining
Sciences 34 (071): 3–4.
88 Li et al.

Espinosa, Horacio D., Pablo D. Zavattieri, and Sunil K Dwivedi. 1998. “A Finite
Deformation Continuum/Discrete Model for the Description of Fragmentation
and Damage in Brittle Materials.” Journal of the Mechanics and Physics of Solids
46 (10): 1909–42. https://doi.org/10.1016/S0022-5096(98)00027-1.
Farbaniec, L., J. D. Hogan, and K. T. Ramesh. 2015. “Micromechanisms Associated
with the Dynamic Compressive Failure of Hot-Pressed Boron Carbide.” Scripta
Materialia 106: 52–56. https://doi.org/10.1016/j.scriptamat.2015.05.004.
Farbaniec, L., J. D. Hogan, K. Y. Xie, M. Shaeffer, K. J. Hemker, and K. T. Ramesh.
2017. “Damage Evolution of Hot-Pressed Boron Carbide under Confined
Dynamic Compression.” International Journal of Impact Engineering 99: 75–84.
https://doi.org/10.1016/j.ijimpeng.2016.09.008.
Farbaniec, Lukasz, James Hogan, James McCauley, and Kaliat Ramesh. 2016.
“Anisotropy of Mechanical Properties in a Hot-Pressed Boron Carbide.”
International Journal of Applied Ceramic Technology 13 (6): 1008–16.
https://doi.org/10.1111/ijac.12585.
Field, J. E., S. M. Walley, W. G. Proud, H. T. Goldrein, and C. R. Siviour. 2004. Review
of Experimental Techniques for High Rate Deformation and Shock Studies.
International Journal of Impact Engineering. Vol. 30.
https://doi.org/10.1016/j.ijimpeng.2004.03.005.
Forrestal, M. J., and D. B. Longcope. 1990. “Target Strength of Ceramic Materials for
High-Velocity Penetration.” Journal of Applied Physics 67 (8): 3669–72.
https://doi.org/10.1063/1.345322.
Forrestal, M J, and D Y Tzou. 1997. “A Spherical Cavity-Expansion Penetration Model
for Concrete Targets.” International Journal of Solids and Structures 34 (31–32):
4127–46. https://doi.org/10.1016/S0020-7683(97)00017-6.
Forrestal, Michael J. 1986. “Penetration into Dry Porous Rock.” International Journal
of Solids and Structures 22 (12): 1485–1500. https://doi.org/10.1016/0020-
7683(86)90057-0.
Fountzoulas, G Costas, and JC LaSalvia. 2013. “Material Models Sensitivity of
Tungsten-Based Penetrators Impacting on Confined Boron-Carbide.” In Dynamic
Behavior of Materials, Volume 1, 251–58. Springer.
Freund, L. B. 1973. “Crack Propagation in an Elastic Solid Subjected to General
Loading-III. Stress Wave Loading.” Journal of the Mechanics and Physics of
Solids 21 (2): 47–61. https://doi.org/10.1016/0022-5096(73)90029-X.
Ge, D., V. Domnich, T. Juliano, E. A. Stach, and Y. Gogotsi. 2004. “Structural Damage
in Boron Carbide under Contact Loading.” Acta Materialia 52 (13): 3921–27.
https://doi.org/10.1016/j.actamat.2004.05.007.
Ghosh, Dipankar, Ghatu Subhash, Chee Huei Lee, and Yoke Khin Yap. 2007. “Strain-
Induced Formation of Carbon and Boron Clusters in Boron Carbide during
Dynamic Indentation.” Applied Physics Letters 91 (6): 1–4.
https://doi.org/10.1063/1.2768316.
Ghosh, Dipankar, Ghatu Subhash, Tirumalai S. Sudarshan, Ramachandran
Radhakrishnan, and Xin Lin Gao. 2007. “Dynamic Indentation Response of Fine-
Grained Boron Carbide.” Journal of the American Ceramic Society 90 (6): 1850–
57. https://doi.org/10.1111/j.1551-2916.2007.01652.x.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 89

Ghosh, Dipankar, Ghatu Subhash, James Q. Zheng, and Virginia Halls. 2012a.
“Influence of Stress State and Strain Rate on Structural Amorphization in Boron
Carbide.” Journal of Applied Physics 111 (6). https://doi.org/10.1063/1.3696971.
———. 2012b. “Influence of Stress State and Strain Rate on Structural Amorphization
in Boron Carbide.” Journal of Applied Physics 111 (6).
https://doi.org/10.1063/1.3696971.
Glenn, L.A., and A. Chudnovskly. 1986. “Strain-Energy Effects on Dynamic
Fragmentation.” Journal of Applied Physics 59 (4): 1379–80.
https://doi.org/10.1063/1.336532.
Gold, Vladimir M, and Ernest L Baker. 2008. “A Model for Fracture of Explosively
Driven Metal Shells.” Engineering Fracture Mechanics 75 (2): 275–89.
https://doi.org/10.1016/j.engfracmech.2007.02.025.
Gong, Daozhen, Stefan Nadolski, Chunbao Sun, Bern Klein, and Jue Kou. 2018. “The
Effect of Strain Rate on Particle Breakage Characteristics.” Powder Technology
339: 595–605. https://doi.org/10.1016/j.powtec.2018.08.020.
Gooch, W. A., M. S. Burkins, G. Hauver, P. Netherwood, and R. Benck. 2000.
“Dynamic X-Ray Imaging of the Penetration of Boron Carbide.” Le Journal de
Physique IV 10 (PR9): Pr9-583-Pr9-588. https://doi.org/10.1051/jp4:2000997.
Gour, Govind, Ahmad Serjouei, and Idapalapati Sridhar. 2017. “Influence of Geometry
and Hardness of the Backing Plate on Ballistic Performance of Bi-Layer Ceramic
Armor.” Procedia Engineering 173: 93–100.
https://doi.org/10.1016/j.proeng.2016.12.040.
Grady, D. E. 1994. “Shock-Wave Strength Properties of Boron Carbide and Silicon
Carbide.” Le Journal de Physique IV 04 (C8): C8-385-C8-391.
https://doi.org/10.1051/jp4:1994859.
Grady, D. 1991. “Shock-Compression Properties of Ceramics.” Albuquerque, NM.
———. 2009. “Dynamic Fragmentation of Solids.” Shock Wave Science and
Technology Reference Library 3: 169–276.
Grady, D E. 2008. “Fragment Size Distributions from the Dynamic Fragmentation of
Brittle Solids.” International Journal of Impact Engineering 35 (12): 1557–62.
https://doi.org/10.1016/j.ijimpeng.2008.07.042.
Grady, DE, and RL Moody. 1996. “Shock Compression Profiles in Ceramics.”
Albuquerque, NM.
Grady, Dennis. 1982. “Local Inertial Effects in Dynamic Fragmentation.” Journal of
Applied Physics 53 (1): 322–25. https://doi.org/10.1063/1.329934.
Grady, Dennis E. 2010. “Length Scales and Size Distributions in Dynamic
Fragmentation.” International Journal of Fracture 163 (1–2): 85–99.
https://doi.org/10.1007/s10704-009-9418-4.
Grady, Dennis, and M.E. Kipp. 1985a. “Geometric Statistics and Dynamic
Fragmentation.” Journal of Applied Physics 58 (3): 1210–22.
https://doi.org/10.1063/1.336139.
———. 1985b. “MECHANISMS OF DYNAMIC FRAGMENTATION: FACTORS
GOVERNING FRAGMENT SIZE.” Mechanics of Materials 4 (3–4): 311–20.
Grady, Dennis, and N.F. Mott. 2006. “A Theory of the Fragmentation of Shells and
Bombs.” Fragmentation of Rings and Shells, no. 24: 243–94.
90 Li et al.

https://doi.org/10.1007/978-3-540-27145-1_11.
Gray III, G.T. 1990. “Shock Experiments in Metals and Ceramics.” New Mexico.
Griggs, David T. 1936. “Deformation of Rocks under High Confining Pressures: I.
Experiments at Room Temperature.” The Journal of Geology 44 (5): 541–77.
Guo, X J, T He, and H M Wen. 2013. “Cylindrical Cavity Expansion Penetration Model
for Concrete Targets with Shear Dilatancy.” Journal of Engineering Mechanics
139 (9): 1260–67. https://doi.org/10.1061/(ASCE)EM.1943-7889.0000547.
Gust, W. H., and E.B. Royce. 1971. “Dynamic Yield Strengths of B4C, BeO, and Al2O
3 Ceramics.” Journal of Applied Physics 42 (1): 276–95.
https://doi.org/10.1063/1.1659584.
Hemker, Kevin J, Mingwei Chen, and James W McCauley. 2003. “Shock-Induced
Localized Amorphization in Boron Carbide.” Science 299 (5612): 1563–66.
https://doi.org/10.1126/science.1080819.
Hogan, James D., Lukasz Farbaniec, Matt Shaeffer, and K. T. Ramesh. 2015. “The
Effects of Microstructure and Confinement on the Compressive Fragmentation of
an Advanced Ceramic.” Journal of the American Ceramic Society 98 (3): 902–
12. https://doi.org/10.1111/jace.13353.
Hogan, James David, Lukasz Farbaniec, Nitin Daphalapurkar, and K. T. Ramesh. 2016.
“On Compressive Brittle Fragmentation.” Journal of the American Ceramic
Society 99 (6): 2159–69. https://doi.org/10.1111/jace.14171.
Hogan, James David, Lukasz Farbaniec, Debjoy Mallick, Vladislav Domnich, Kanak
Kuwelkar, Tomoko Sano, James W. McCauley, and Kaliat T. Ramesh. 2017.
“Fragmentation of an Advanced Ceramic under Ballistic Impact: Mechanisms
and Microstructure.” International Journal of Impact Engineering 102: 47–54.
https://doi.org/10.1016/j.ijimpeng.2016.12.008.
Hogan, James David, Lukasz Farbaniec, Tomoko Sano, Matthew Shaeffer, and K. T.
Ramesh. 2016. “The Effects of Defects on the Uniaxial Compressive Strength
and Failure of an Advanced Ceramic.” Acta Materialia 102: 263–72.
https://doi.org/10.1016/j.actamat.2015.09.028.
Holmquist, Gordon R, and TJ Johnson. 1999. “Response of Boron Carbide Subjected
to Large Strains , High Strain Rates , and High Pressures Response of Boron
Carbide Subjected to Large Strains , High Strain Rates , and High Pressures.”
Journal of Applied Physics 85 (12): 8060--8073.
https://doi.org/10.1063/1.370643.
Holmquist, T. J., and G. R. Johnson. 2006. “Characterization and Evaluation of Boron
Carbide for Plate-Impact Conditions.” Journal of Applied Physics 100 (9).
https://doi.org/10.1063/1.2362979.
———. 2012. “Incorporation of the Deshpande-Evans Mechanism-Based Damage
Model into the EPIC Code.” AIP Conference Proceedings 1426 (March): 68–71.
https://doi.org/10.1063/1.3686223.
Holmquist, Timothy J., and Gordon R. Johnson. 2011. “A Computational Constitutive
Model for Glass Subjected to Large Strains, High Strain Rates and High
Pressures.” Journal of Applied Mechanics 78 (5): 051003.
https://doi.org/10.1115/1.4004326.
Holmquist, Timothy J, and Gordon R Johnson. 2005. “Modeling Prestressed Ceramic
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 91

and Its Effect on Ballistic Performance.” International Journal of Impact


Engineering 31 (2): 113–27. https://doi.org/10.1016/j.ijimpeng.2003.11.002.
Horii, H., and S. Nemat-Nasser. 1983. “Overall Moduli of Solids with Microcracks:
Load-Induced Anisotropy.” Journal of the Mechanics and Physics of Solids 31
(2): 155–71. https://doi.org/10.1016/0022-5096(83)90048-0.
———. 1985. “Compression-Induced Microcrack Growth in Brittle Solids: Axial
Splitting and Shear Failure.” Journal of Geophysical Research 90 (B4): 3105.
https://doi.org/10.1029/jb090ib04p03105.
Hosson, Jeff Th M De. 2009. “Advances in Transmission Electron Microscopy: In Situ
Straining and in Situ Compression Experiments on Metallic Glasses.”
Microscopy Research and Technique 72 (3): 250–60.
https://doi.org/10.1002/jemt.20678.
Hu, Guangli, Junwei Liu, Lori Graham-Brady, and K. T. Ramesh. 2015. “A 3D
Mechanistic Model for Brittle Materials Containing Evolving Flaw Distributions
under Dynamic Multiaxial Loading.” Journal of the Mechanics and Physics of
Solids 78: 269–97. https://doi.org/10.1016/j.jmps.2015.02.014.
Johnson, Gordon R., and Tim J. Holmquist. 2008. “An Improved Computational
Constitutive Model for Brittle Materials” 981 (May): 981–84.
https://doi.org/10.1063/1.46199.
Johnson, Gordon R, and Tim J Holmquist. 1994. “An Improved Computational
Constitutive Model for Brittle Materials.” In AIP Conference Proceedings, 981–
84. AIP.
Johnson, GR, and TJ Holmquist. 1992. “A Computational Constitutive Model for
Brittle Materials Subjected to Large Strains, High Strain Rates and High
Pressures.” Shock Wave and High-Strain-Rate Phenomena in Materials, 1075--
1081.
Ju, By J W, X Lee, and Associate Member. 1991. “Micromechanical Damage Models
for Brittle Solids I: Tensile Loading” 117 (7): 1495–1514.
Kiener, D, and A M Minor. 2011. “Source Truncation and Exhaustion: Insights from
Quantitative in Situ TEM Tensile Testing.” Nano Letters 11 (9): 3816–20.
https://doi.org/10.1021/nl201890s.
Kolari, Kari. 2017. “A Complete Three-Dimensional Continuum Model of Wing-Crack
Growth in Granular Brittle Solids.” International Journal of Solids and Structures
115–116: 27–42. https://doi.org/10.1016/j.ijsolstr.2017.02.012.
Kolsky, Herbert. 1963. Stress Waves in Solids. Courier Corporation.
Korotaev, Pavel, Pavel Pokatashkin, and Aleksey Yanilkin. 2016. “The Role of Non-
Hydrostatic Stresses in Phase Transitions in Boron Carbide.” Computational
Materials Science 121: 106–12.
https://doi.org/10.1016/j.commatsci.2016.04.041.
Kraft, R H, J F Molinari, K T Ramesh, and D H Warner. 2008. “Computational
Micromechanics of Dynamic Compressive Loading of a Brittle Polycrystalline
Material Using a Distribution of Grain Boundary Properties.” Journal of the
Mechanics and Physics of Solids 56 (8): 2618–41.
https://doi.org/10.1016/j.jmps.2008.03.009.
Krell, Andreas, and Elmar Strassburger. 2014. “Order of Influences on the Ballistic
92 Li et al.

Resistance of Armor Ceramics and Single Crystals.” Materials Science and


Engineering A 597: 422–30. https://doi.org/10.1016/j.msea.2013.12.101.
Krimsky, Erez, K T Ramesh, M Bratcher, M Foster, and James David Hogan. 2019.
“Quantification of Damage and Its Effects on the Compressive Strength of an
Advanced Ceramic.” Engineering Fracture Mechanics 208 (September 2018):
107–18. https://doi.org/10.1016/j.engfracmech.2019.01.007.
Krishnan, K, S Sockalingam, S Bansal, and S D Rajan. 2010. “Composites : Part B
Numerical Simulation of Ceramic Composite Armor Subjected to Ballistic
Impact.” Composites Part B 41 (8): 583–93.
https://doi.org/10.1016/j.compositesb.2010.10.001.
Lankford, J, W. W. Predebon, J. M. Staehler, G. Subhash, B. J. Pletka, and C. E.
Anderson. 1998. “The Role of Plasticity as a Limiting Factor in the Compressive
Failure of High Strength Ceramics.” Mechanics of Materials 29 (3–4): 205–18.
https://doi.org/10.1016/S0167-6636(98)00023-4.
LaSalvia, JC, RB Leavy, JR Houskamp, HT Miller, DE MacKenzie, and J Campbell.
2009. “Ballistic Impact Damage Observations in a Hot-Pressed Boron Carbide.”
In Ceramic Engineering and Science Proceedings, 45.
LaSalvia, JC, RC McCuiston, G Fanchini, JW McCauley, M Chhowalla, HT Miller,
and DE MacKenzie. 2007. “Shear Localization in a Sphere-Impacted Armor-
Grade Boron Carbide.” In Proceedings of the 23rd International Symposium on
Ballistics, 1329–37.
LaSalvia, JC, MJ Normandia, HT Miller, and DE MacKenzie. 2009. “Sphere Impact
Induced Damage in Ceramics: II. Armor-Grade B4C and WC.” In Advances in
Ceramic Armor: A Collection of Papers Presented at the 29th International
Conference on Advanced Ceramics and Composites, Jan 23-28, 2005, Cocoa
Beach, FL, 183. John Wiley & Sons.
Leong, Andrew FT, Andrew K Robinson, K Fezzaa, T Sun, N Sinclair, DT Casem, PK
Lambert, et al. 2018. “Quantitative In Situ Studies of Dynamic Fracture in Brittle
Solids Using Dynamic X-Ray Phase Contrast Imaging.” Experimental Mechanics
58 (9): 1423–37. https://doi.org/10.1007/s11340-018-0414-3.
Levy, S., and J. F. Molinari. 2010. “Dynamic Fragmentation of Ceramics, Signature of
Defects and Scaling of Fragment Sizes.” Journal of the Mechanics and Physics
of Solids 58 (1): 12–26. https://doi.org/10.1016/j.jmps.2009.09.002.
Li, H. Y., P. Motamedi, and J. D. Hogan. 2019. “Characterization and Mechanical
Testing on Novel (Γ + α 2 ) – TiAl/Ti 3 Al/Al 2 O 3 Cermet.” Materials Science
and Engineering A 750 (January): 152–63.
https://doi.org/10.1016/j.msea.2019.02.039.
Li, X. F., X. Li, H B Li, Q B Zhang, and J Zhao. 2018. “Dynamic Tensile Behaviours
of Heterogeneous Rocks: The Grain Scale Fracturing Characteristics on Strength
and Fragmentation.” International Journal of Impact Engineering 118 (April):
98–118. https://doi.org/10.1016/j.ijimpeng.2018.04.006.
Li, X F, Q. B. Zhang, H. B. Li, and J. Zhao. 2018. “Grain-Based Discrete Element
Method (GB-DEM) Modelling of Multi-Scale Fracturing in Rocks Under
Dynamic Loading.” Rock Mechanics and Rock Engineering 51 (12): 3785–3817.
https://doi.org/10.1007/s00603-018-1566-2.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 93

Liu, Yue, Tianlu Zhao, Wangwei Ju, Siqi Shi, Siqi Shi, and Siqi Shi. 2017. “Materials
Discovery and Design Using Machine Learning.” Journal of Materiomics.
Elsevier Taiwan LLC. https://doi.org/10.1016/j.jmat.2017.08.002.
Maiti, Spandan, Krishnan Rangaswamy, and Philippe H Geubelle. 2005. “Mesoscale
Analysis of Dynamic Fragmentation of Ceramics under Tension.” Acta
Materialia 53 (3): 823–34. https://doi.org/10.1016/j.actamat.2004.10.034.
Mallick, D D, M Zhao, J Parker, V Kannan, B T Bosworth, D Sagapuram, M A Foster,
and K T Ramesh. 2019. “Laser-Driven Flyers and Nanosecond-Resolved
Velocimetry for Spall Studies in Thin Metal Foils.” Experimental Mechanics.
https://doi.org/10.1007/s11340-019-00519-x.
McCauley, J. W., E. Strassburger, P. Patel, B. Paliwal, and K. T. Ramesh. 2013.
“Experimental Observations on Dynamic Response of Selected Transparent
Armor Materials.” Experimental Mechanics 53 (1): 3–29.
https://doi.org/10.1007/s11340-012-9658-5.
Meyer, L, I Faber, L Meyer, I Faber Investigations, Ceramic Powder, and Journal De
Physique. 1997. “Investigations on Granular Ceramics and Ceramic Powder.”
Miller, O, L B Freund, and A Needleman. 1999. “Modeling and Simulation of Dynamic
Fragmentation in Brittle Materials.” International Journal of Fracture 96 (2):
101–25. https://doi.org/10.1023/A:1018666317448.
Moller, Jesper. 1992. “Random Johnson-Mehl Tessellations.” Advances in Applied
Probability 24 (4): 814–44.
Mott, N.F. 1945. “Fragmentation of Shell Cases.” In Proceedings of the Royal Society
of London. Series A. Mathematical and Physical Sciences, 300–308. The Royal
Society London.
Mott, NF, and E Linfoot. 1943. “A Theory of Fragmentation.”
Moynihan, T.J., J.C. LaSalvia, and M.S. Burkins. 2002. “Analysis of Shatter Gap
Phenomenon in a Boron Carbide/Composite Laminate Armor System.” In 20th
International Ballistics Symposium, edited by Orphal D Carleone J, 1096–1103.
Lancaster, PA: DEStech Publications.
Nemat-Nasser, S., and H. Horii. 1982. “Compression-Induced Nonplanar Crack
Extension with Application to Splitting, Exfoliation, and Rockburst.” Journal of
Geophysical Research 87 (B8): 6805–21.
https://doi.org/10.1029/JB087iB08p06805.
Nemat-Nasser, S., and M. Obata. 1988. “A Microcrack Model of Dilatancy in Brittle
Materials.” Journal of Applied Mechanics 55 (1): 24.
https://doi.org/10.1115/1.3173647.
Orphal, D.L., R.R. Franzen, A.C. Charters, T.L. Menna, and A.J. Piekutowski. 1997.
“Penetration of Confined Boron Carbide Targets by Tungsten Long Rods at
Impact Velocities from 1.5 to 5.0 Km/S.” International Journal of Impact
Engineering 19 (1): 15–29. https://doi.org/10.1016/s0734-743x(96)00004-8.
Pack, DC, and WM Evans. 1951. “Penetration by High-Velocity (`Munroe’) Jets: I.” In
Proceedings of the Physical Society. Section B, 298. IOP Publishing.
Paliwal, B., and K. T. Ramesh. 2007. “Effect of Crack Growth Dynamics on the Rate-
Sensitive Behavior of Hot-Pressed Boron Carbide.” Scripta Materialia 57 (6):
481–84. https://doi.org/10.1016/j.scriptamat.2007.05.028.
94 Li et al.

———. 2008. “An Interacting Micro-Crack Damage Model for Failure of Brittle
Materials under Compression.” Journal of the Mechanics and Physics of Solids
56 (3): 896–923. https://doi.org/10.1016/j.jmps.2007.06.012.
Paliwal, B., K. T. Ramesh, and J. W. McCauley. 2006. “Direct Observation of the
Dynamic Compressive Failure of a Transparent Polycrystalline Ceramic
(AION).” Journal of the American Ceramic Society 89 (7): 2128–33.
https://doi.org/10.1111/j.1551-2916.2006.00965.x.
Palomares-García, Alberto Jesús, Maria Teresa Pérez-Prado, and Jon Mikel Molina-
Aldareguia. 2017. “Effect of Lamellar Orientation on the Strength and Operating
Deformation Mechanisms of Fully Lamellar TiAl Alloys Determined by
Micropillar Compression.” Acta Materialia 123: 102–14.
https://doi.org/10.1016/j.actamat.2016.10.034.
Paris, V., N. Frage, M. P. Dariel, and E. Zaretsky. 2010. “The Spall Strength of Silicon
Carbide and Boron Carbide Ceramics Processed by Spark Plasma Sintering.”
International Journal of Impact Engineering 37 (11): 1092–99.
https://doi.org/10.1016/j.ijimpeng.2010.06.008.
Parsard, Gregory, Ghatu Subhash, and Phillip Jannotti. 2018. “Amorphization-Induced
Volume Change and Residual Stresses in Boron Carbide.” Journal of the
American Ceramic Society 101 (6): 2606–15. https://doi.org/10.1111/jace.15417.
Pensée, V., D. Kondo, and L. Dormieux. 2002. “Micromechanical Analysis of
Anisotropic Damage in Brittle Materials.” Journal of Engineering Mechanics 128
(8): 889–97. https://doi.org/10.1061/(asce)0733-9399(2002)128:8(889).
Raiser, G., R. J. Clifton, and M. Ortiz. 1990. “A Soft-Recovery Plate Impact Experiment
for Studying Microcracking in Ceramics.” Mechanics of Materials 10 (1–2): 43–
58. https://doi.org/10.1016/0167-6636(90)90016-9.
Raiser, G.F., J.L. Wise, R. J. Clifton, Dennis Grady, and D.E. Cox. 1994. “Plate Impact
Response of Ceramics and Glasses.” Journal of Applied Physics 75 (8): 3862–69.
https://doi.org/10.1063/1.356066.
Rajendran, A. M., and D. J. Grove. 1996. “Modeling the Shock Response of Silicon
Carbide, Boron Carbide and Titanium Diboride.” International Journal of Impact
Engineering 18 (6): 611–31. https://doi.org/10.1016/0734-743X(96)89122-6.
Rajendran, A. M., and J. L. Kroupa. 1989. “Impact Damage Model for Ceramic
Materials.” Journal of Applied Physics 66 (8): 3560–65.
https://doi.org/10.1063/1.344085.
Rajendran, AM. 1994. “Modeling the Impact Behavior of AD85 Ceramic Under Multi-
Axial Loading".” International Journal of Impact Engineering 15 (6): 749–68.
Ramesh, K T, James D Hogan, Jamie Kimberley, and Angela Stickle. 2015. “A Review
of Mechanisms and Models for Dynamic Failure, Strength, and Fragmentation.”
In Planetary and Space Science, 107:10–23. Elsevier.
https://doi.org/10.1016/j.pss.2014.11.010.
Ravichandran, G., and G. Subhash. 1995. “A Micromechanical Model for High Strain
Rate Behavior of Ceramics.” International Journal of Solids and Structures 32
(17–18): 2627–46. https://doi.org/10.1016/0020-7683(94)00286-6.
Reddy, K. Madhav, P. Liu, A. Hirata, T. Fujita, and M. W. Chen. 2013. “Atomic
Structure of Amorphous Shear Bands in Boron Carbide.” Nature
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 95

Communications 4: 1–5. https://doi.org/10.1038/ncomms3483.


Renzo, Alberto Di, Francesco Paolo, and Di Maio. 2004. “Comparison of Contact-Force
Models for the Simulation of Collisions in DEM-Based Granular Ow Codes” 59:
525–41. https://doi.org/10.1016/j.ces.2003.09.037.
Romeis, Stefan, Jonas Paul, M Ziener, and Wolfgang Peukert. 2012. “A Novel
Apparatus for in Situ Compression of Submicron Structures and Particles in a
High Resolution SEM.” Review of Scientific Instruments 83 (9).
https://doi.org/10.1063/1.4749256.
Rosenberg, Z, NS Brar, and S. J. Bless. 1991. “Dynamic High-Pressure Properties of
AlN Ceramic as Determined by Flyer Plate Impact.” Journal of Applied Physics
70 (1): 167–71. https://doi.org/10.1063/1.350337.
Roy, T. K., C. Subramanian, and A. K. Suri. 2006. “Pressureless Sintering of Boron
Carbide.” Ceramics International 32 (3): 227–33.
https://doi.org/10.1016/j.ceramint.2005.02.008.
Sammis, C. G., and M. F. Ashby. 1986. “The Failure of Brittle Porous Solids under
Compressive Stress States.” Acta Metallurgica 34 (3): 511–26.
https://doi.org/10.1016/0001-6160(86)90087-8.
Sano, Tomoko, Matthew Shaeffer, Lionel Vargas-gonzalez, and Joshua Pomerantz.
2018. “Dynamic Behavior of Materials, Volume 1” 1: 13–19.
https://doi.org/10.1007/978-3-319-62956-8.
Satapathy, Sikhanda. 2001. “Dynamic Spherical Cavity Expansion in Brittle Ceramics.”
International Journal of Solids and Structures 38 (32–33): 5833–45.
https://doi.org/10.1016/S0020-7683(00)00388-7.
Satapathy, Sikhanda S., and Stephan Bless. 1996. “Calculation of Penetration
Resistance of Brittle Materials Using Spherical Cavity Expansion Analysis.”
Mechanics of Materials 23 (4): 323–30. https://doi.org/10.1016/0167-
6636(96)00022-1.
Satapathy, Sikhanda S., and Stephan J. Bless. 2000. “Cavity Expansion Resistance of
Brittle Materials Obeying a Two-Curve Pressure-Shear Behavior.” Journal of
Applied Physics 88 (7): 4004–12. https://doi.org/10.1063/1.1288007.
Savio, S. G., K. Ramanjaneyulu, V. Madhu, and T. Balakrishna Bhat. 2011. “An
Experimental Study on Ballistic Performance of Boron Carbide Tiles.”
International Journal of Impact Engineering 38 (7): 535–41.
https://doi.org/10.1016/j.ijimpeng.2011.01.006.
Shafiq, M, and G Subhash. 2016a. “An Extended Mohr-Coulomb Model for Fracture
Strength of Intact Brittle Materials under Ultrahigh Pressures.” Journal of the
American Ceramic Society 99 (2): 627–30. https://doi.org/10.1111/jace.14026.
———. 2016b. “Dynamic Deformation Characteristics of Zirconium Diboride-Silicon
Carbide under Multi-Axial Confinement.” International Journal of Impact
Engineering 91: 158–69. https://doi.org/10.1016/j.ijimpeng.2016.01.009.
Shenoy, V B, and K. S. Kim. 2003. “Disorder Effects in Dynamic Fragmentation of
Brittle Materials.” In Journal of the Mechanics and Physics of Solids, 51:2023–
35. https://doi.org/10.1016/j.jmps.2003.09.010.
Shockey, Donald A., A. H. Marchand, S. R. Skaggs, G. E. Cort, M. W. Burkett, and R.
Parker. 1990. “Failure Phenomenology of Confined Ceramic Targets and
96 Li et al.

Impacting Rods.” International Journal of Impact Engineering 9 (3): 263–75.


https://doi.org/10.1016/0734-743X(90)90002-D.
Shokrieh, M M, and G H Javadpour. 2008. “Penetration Analysis of a Projectile in
Ceramic Composite Armor.” Composite Structures 82 (2): 269–76.
https://doi.org/10.1016/j.compstruct.2007.01.023.
Simha, C. Hari Manoj, S. J. Bless, and A. Bedford. 2001. “Computational Modeling of
the Penetration Response of a High-Purity Ceramic.” International Journal of
Impact Engineering 27 (1): 65–86. https://doi.org/10.1016/S0734-
743X(01)00036-7.
Song, Hong Wei, Qi Jian He, Ji Jia Xie, and A Tobota. 2008. “Fracture Mechanisms
and Size Effects of Brittle Metallic Foams: In Situ Compression Tests inside
SEM.” Composites Science and Technology 68 (12): 2441–50.
https://doi.org/10.1016/j.compscitech.2008.04.023.
Strassburger, E. 2014. “Edge-on Impact Investigation of Fracture Propagation in Boron
Carbide.” In Advances in Ceramic Armor IX . Edited by Jerry C . LaSalvia ., 15–
24. John Wiley & Sons, Inc.
Subhash, Ghatu, Amnaya P. Awasthi, Cody Kunka, Phillip Jannotti, and Matthew
DeVries. 2016. “In Search of Amorphization-Resistant Boron Carbide.” Scripta
Materialia 123: 158–62. https://doi.org/10.1016/j.scriptamat.2016.06.012.
Subhash, Ghatu, Dipankar Ghosh, Justin Blaber, James Q. Zheng, Virginia Halls, and
Karl Masters. 2013. “Characterization of the 3-D Amorphized Zone beneath a
Vickers Indentation in Boron Carbide Using Raman Spectroscopy.” Acta
Materialia 61 (10): 3888–96. https://doi.org/10.1016/j.actamat.2013.03.028.
Subhash, Ghatu, Dipankar Ghosh, and Spandan Maiti. 2009. “Static and Dynamic
Indentation Response of Fine Grained Boron Carbide.” In Advances in Ceramic
Armor III, 29–44. https://doi.org/10.1002/9780470339695.ch3.
Subhash, Ghatu, Spandan Maiti, Philippe H. Geubelle, and Dipankar Ghosh. 2008.
“Recent Advances in Dynamic Indentation Fracture, Impact Damage and
Fragmentation of Ceramics.” Journal of the American Ceramic Society 91 (9):
2777–91. https://doi.org/10.1111/j.1551-2916.2008.02624.x.
Swab, Jeffrey J., Christopher S. Meredith, Daniel T. Casem, and William Robert
Gamble. 2017. “Static and Dynamic Compression Strength of Hot-Pressed Boron
Carbide Using a Dumbbell-Shaped Specimen.” Journal of Materials Science 52
(17): 10073–84. https://doi.org/10.1007/s10853-017-1210-7.
Tapponnier, Paul, and WF Brace. 1976. “Development of Stress-Induced Microcrakcs
in Westerly Granite.” In International Journal of Rock Mechanics and Mining
Sciences & Geomechanics Abstracts, 13:103–12. Elsevier.
Tate, A. 1967. “A Theory for the Deceleration of Long Rods after Impact.” Journal of
the Mechanics and Physics of Solids 15 (6): 387–99.
https://doi.org/10.1016/0022-5096(67)90010-5.
Tate, A. 1969. “Further Results in the Theory of Long Rod Penetration.” Journal of the
Mechanics and Physics of Solids 17 (3): 141–50.
———. 1986. “Long Rod Penetration Models-Part II. Extensions to the Hydrodynamic
Theory of Penetration.” International Journal of Mechanical Sciences 28 (9):
599–612. https://doi.org/10.1016/0020-7403(86)90075-5.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 97

Taylor, De Carlos E. 2015. “Shock Compression of Boron Carbide: A Quantum


Mechanical Analysis.” Journal of the American Ceramic Society 98 (10): 3308–
18. https://doi.org/10.1111/jace.13711.
Taylor, Decarlos E., James W. McCauley, and T. W. Wright. 2012. “The Effects of
Stoichiometry on the Mechanical Properties of Icosahedral Boron Carbide under
Loading.” Journal of Physics Condensed Matter 24 (50).
https://doi.org/10.1088/0953-8984/24/50/505402.
Tekalur, Srinivasan Arjun, and Oishik Sen. 2011. “Effect of Specimen Size in the
Kolsky Bar.” Procedia Engineering 10: 2663–71.
https://doi.org/10.1016/j.proeng.2011.04.444.
Terzaghi, Karl, Ralph B Peck, and Gholamreza Mesri. 1996. Soil Mechanics in
Engineering Practice. John Wiley & Sons, Inc.
Tonge, Andrew L., and K. T. Ramesh. 2016a. “Multi-Scale Defect Interactions in High-
Rate Brittle Material Failure. Part I: Model Formulation and Application to
ALON.” Journal of the Mechanics and Physics of Solids 86: 117–49.
https://doi.org/10.1016/j.jmps.2015.10.007.
———. 2016b. “Multi-Scale Defect Interactions in High-Rate Brittle Material Failure.
Part I: Model Formulation and Application to ALON.” Journal of the Mechanics
and Physics of Solids 86: 117–49. https://doi.org/10.1016/j.jmps.2015.10.007.
———. 2016c. “Multi-Scale Defect Interactions in High-Rate Failure of Brittle
Materials, Part II: Application to Design of Protection Materials.” Journal of the
Mechanics and Physics of Solids 86: 237–58.
https://doi.org/10.1016/j.jmps.2015.10.006.
Tonge, Andrew L., and Brian E Schuster. 2017. “Towards Predictive Transferable
Simulations of Ceramic Failure in Ballistic Events.” In Procedia Engineering,
204:247–54. Elsevier B.V. https://doi.org/10.1016/j.proeng.2017.09.732.
Toussaint, G, and Ioannis Polyzois. 2019. “Steel Spheres Impact on Alumina Ceramic
Tiles: Experiments and Finite Element Simulations.” International Journal of
Applied Ceramic Technology.
Tracy, S J, R F Smith, J K Wicks, D E Fratanduono, A E Gleason, C A Bolme, V B
Prakapenka, et al. 2019. “Insitu Observation of a Phase Transition in Silicon
Carbide under Shock Compression Using Pulsed X-Ray Diffraction.” Physical
Review B 99 (21): 1–7. https://doi.org/10.1103/physrevb.99.214106.
Turcotte, Donald L. 1997. Fractals and Chaos in Geology and Geophysics. Cambridge
university press.
Vargas-Gonzalez, Lionel, Robert F. Speyer, and James Campbell. 2010. “Flexural
Strength, Fracture Toughness, and Hardness of Silicon Carbide and Boron
Carbide Armor Ceramics.” International Journal of Applied Ceramic Technology
7 (5): 643–51. https://doi.org/10.1111/j.1744-7402.2010.02501.x.
Viggiani, G, N Lenoir, M Bornert, J Desrues, and P Be. 2007. “Volumetric Digital
Image Correlation Applied to X-Ray Microtomography Images from Triaxial
Compression Tests on Argillaceous Rock.” Strain 43: 193–205.
Vogler, T. J., W. D. Reinhart, and L. C. Chhabildas. 2004. “Dynamic Behavior of Boron
Carbide.” Journal of Applied Physics 95 (8): 4173–83.
https://doi.org/10.1063/1.1686902.
98 Li et al.

Walker, James D. 2003. “Analytically Modeling Hypervelocity Penetration of Thick


Ceramic Targets.” International Journal of Impact Engineering 29 (1–10): 747–
55.
Walker, James D, and Charles E Anderson Jr. 1995. “A Time-Dependent Model for
Long-Rod Penetration.” International Journal of Impact Engineering 16 (1): 19–
48.
Wang, Jinghan, Ju Li, Sidney Yip, Simon Phillpot, and Dieter Wolf. 1995. “Mechanical
Instabilities of Homogeneous Crystals.” Physical Review B 52 (17): 12627–35.
https://doi.org/10.1103/PhysRevB.52.12627.
Wang, Zhiyong, and Peifeng Li. 2015. “Dynamic Failure and Fracture Mechanism in
Alumina Ceramics: Experimental Observations and Finite Element Modelling.”
Ceramics International 41 (10): 12763–72.
https://doi.org/10.1016/j.ceramint.2015.06.110.
Ward, Logan, Stephanie C. O’Keeffe, Joseph Stevick, Glenton R Jelbert, Muratahan
Aykol, and Chris Wolverton. 2018. “A Machine Learning Approach for
Engineering Bulk Metallic Glass Alloys.” Acta Materialia 159: 102–11.
https://doi.org/10.1016/j.actamat.2018.08.002.
Westerling, L., P. Lundberg, and B. Lundberg. 2001. “Tungsten Long-Rod Penetration
into Confined Cylinders of Boron Carbide at and above Ordnance Velocities.”
International Journal of Impact Engineering 25 (7): 703–14.
https://doi.org/10.1016/S0734-743X(00)00072-5.
Wilkins, M L. 1968. “Third Progress Report of Light Armour Program.”
Winkler, Wolf-Dieter, and Alois J. Stilp. 1992. SPALLATION BEHAVIOR OF TiB2,
SiC, AND B4C UNDER PLANAR IMPACT TENSILE STRESSES. Shock
Compression of Condensed Matter–1991. Elsevier B.V.
https://doi.org/10.1016/b978-0-444-89732-9.50108-4.
Woodward, R. L., W. A. Gooch, R. G. O’Donnell, W. J. Perciballi, B. J. Baxter, and S.
D. Pattie. 1994. “A Study of Fragmentation in the Ballistic Impact of Ceramics.”
International Journal of Impact Engineering 15 (5): 605–18.
https://doi.org/10.1016/0734-743X(94)90122-2.
Yan, X. Q., Z. Tang, L. Zhang, J. J. Guo, C. Q. Jin, Y. Zhang, T. Goto, J. W. McCauley,
and M. W. Chen. 2009. “Depressurization Amorphization of Single-Crystal
Boron Carbide.” Physical Review Letters 102 (7): 1–4.
https://doi.org/10.1103/PhysRevLett.102.075505.
Ye, Jia, Raja K Mishra, Anil K. Sachdev, and Andrew M Minor. 2011. “In Situ TEM
Compression Testing of Mg and Mg-0.2 Wt.% Ce Single Crystals.” Scripta
Materialia 64 (3): 292–95. https://doi.org/10.1016/j.scriptamat.2010.09.047.
Zaera, R., and V. Sánchez-Gálvez. 1998. “Analytical Modelling of Normal and Oblique
Ballistic Impact on Ceramic/Metal Lightweight Armours.” International Journal
of Impact Engineering 21 (3): 133–48. https://doi.org/10.1016/S0734-
743X(97)00035-3.
Zavattieri, P D, P V Raghuram, and H D Espinosa. 2001. “Computational Model of
Ceramic Microstructures Subjected to Multi-Axial Dynamic Loading.” Journal
of the Mechanics and Physics of Solids 49 (1): 27–68.
https://doi.org/10.1016/S0022-5096(00)00028-4.
A Review of Dynamic Fracture and Fragmentation of Boron Carbide 99

Zavattieri, Pablo D., and Horacio D Espinosa. 2001. “Grain Level Analysis of Crack
Initiation And.” Acta Materialia 4 (20): 4291–4311.
http://www.sciencedirect.com/science/article/pii/S1359645401002920.
Zeng, Qinglei, Andrew L. Tonge, and K.T. Ramesh. 2019. “A Multi-Mechanism
Constitutive Model for the Dynamic Failure of Quasi-Brittle Materials. Part I-
Amorphization as a Failure Mode.” Journal of the Mechanics and Physics of
Solids 130: 370–92. https://doi.org/10.1016/j.jmps.2019.06.012.
Zhang, Y., T. Mashimo, Y. Uemura, M. Uchino, M. Kodama, K. Shibata, K. Fukuoka,
M. Kikuchi, T. Kobayashi, and T. Sekine. 2006. “Shock Compression Behaviors
of Boron Carbide (B4C).” Journal of Applied Physics 100 (11): 1–6.
https://doi.org/10.1063/1.2399334.
Zhang, Ying, and Chen Ling. 2018. “A Strategy to Apply Machine Learning to Small
Datasets in Materials Science.” Npj Computational Materials 4 (1): 28–33.
https://doi.org/10.1038/s41524-018-0081-z.
Zhao, Shiteng, Bimal Kad, Bruce A. Remington, Jerry C. LaSalvia, Christopher E.
Wehrenberg, Kristopher D. Behler, and Marc A. Meyers. 2016. “Directional
Amorphization of Boron Carbide Subjected to Laser Shock Compression.”
Proceedings of the National Academy of Sciences 113 (43): 12088–93.
https://doi.org/10.1073/pnas.1604613113.
Zhou, Fenghua, Jean François Molinari, and K. T. Ramesh. 2005. “A Cohesive Model
Based Fragmentation Analysis: Effects of Strain Rate and Initial Defects
Distribution.” International Journal of Solids and Structures 42 (18–19): 5181–
5207. https://doi.org/10.1016/j.ijsolstr.2005.02.009.
———. 2006a. “Analysis of the Brittle Fragmentation of an Expanding Ring.”
Computational Materials Science 37 (1–2): 74–85.
https://doi.org/10.1016/j.commatsci.2005.12.017.
———. 2006b. “Characteristic Fragment Size Distributions in Dynamic
Fragmentation.” Applied Physics Letters 88 (26): 16–19.
https://doi.org/10.1063/1.2216892.
Zhu, Wan Cheng, and C. A. Tang. 2004. “Micromechanical Model for Simulating the
Fracture Process of Rock.” Rock Mechanics and Rock Engineering 37 (1): 25–
56. https://doi.org/10.1007/s00603-003-0014-z.
Zoback, Mark D., and James D. Byerlee. 1975. “The Effect of Microcrack Dilatancy on
the Permeability of Westerly Granite.” Journal of Geophysical Research 80 (5):
752–55. https://doi.org/10.1029/jb080i005p00752.
Zuo, Q. H., F. L. Addessio, J. K. Dienes, and M. W. Lewis. 2006. “A Rate-Dependent
Damage Model for Brittle Materials Based on the Dominant Crack.”
International Journal of Solids and Structures 43 (11–12): 3350–80.
https://doi.org/10.1016/j.ijsolstr.2005.06.083.
Zuo, Q. H., D. Disilvestro, and J. D. Richter. 2010. “A Crack-Mechanics Based Model
for Damage and Plasticity of Brittle Materials under Dynamic Loading.”
International Journal of Solids and Structures 47 (20): 2790–98.
https://doi.org/10.1016/j.ijsolstr.2010.06.009.

View publication stats

You might also like