Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Final Scientific Report

For Project Titled:

High-pressure polymorphism of two high-strength ceramics:


Boron carbide (B4C) and silicon carbide (SiC)

Prepared For
United State Department of Energy
Office of Fusion Energy Sciences

Award: DE-SC0016242

Performance Period: September 1, 2016 – August 31, 2017

Prepared by:

Professor Thomas S. Duffy


Department of Geosciences and
Princeton Institute for the Science and Technology of Materials
Princeton University

Submitted: October 31, 2017

1
Disclaimer

Any findings, opinions, and conclusions expressed in this report are those of the author and
do not necessarily reflect the views of the Department of Energy.

Acknowledgements

We thank the staff of the Materials in Extreme Conditions Beamline at the Linac Coherent
Light Source for experimental assistance. Use of the Linac Coherent Light Source, SLAC
National Accelerator Laboratory, is supported by the U.S. Department of Energy, Office of
Science, and Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515.
This report is based upon work supported by the U. S. Department of Energy under award
No DE- SC0016242.

Proprietary Data Notice

This report does not contain any proprietary data.

2
Executive Summary

We have determined the lattice-level structure of two high-strength, technologically


important ceramic materials under shock wave compression. For decades, laboratory shock wave
studies have provided crucial information on the dynamic response of many types of materials.
However, previous studies were limited to continuum-level information only. By combining
laser compression with pulsed X-ray diffraction we are able to obtain in situ structural
information on these materials for the first time.
Silicon carbide (SiC) and boron carbide (B4C) are two high-strength ceramics with many
technical applications including debris shielding, armoring, and coating. We have carried out X-
ray diffraction measurements of the structure of SiC and B4C under shock compression to 185
and 300 GPa, respectively. Experiments were performed at the Matter in Extreme Conditions
beamline of the Linac Coherent Light Source (LCLS). At LCLS, shock waves are driven into the
sample using a 40-J laser with a ~12-15 ns pulse length. The sample is probed with X-rays from
the LCLS free electron laser providing 1012 photons in a monochromatic pulse at 8.5 keV
energy. Diffraction is recorded using pixel array detectors. By varying the delay between the
laser and the x-ray probe, the sample can be studied at various times relative to the shock wave
entering the sample. The diffraction data has been analyzed to determine the crystal structures,
lattice parameters, density, and crystallographic texture as a function of pressure and loading
time.

Our major findings are as follows:

1) Silicon Carbide

• SiC transforms from its low-pressure tetrahedrally coordinated structure to the


octahedrally coordinated rocksalt (B1) phase at ~120 GPa on nanosecond time
scales. This result is consistent with the transition inferred from gas-gun data
despite largely different time scales. There is no evidence for intermediate
structures predicted theoretically.

• Densities determined from X-ray diffraction are in agreement with continuum


measurements and theoretical calculations.

• Upon release, the B1 phase reverts to the low-pressure tetrahedral structure on


nanosecond timescales. At late times, we observe a mixture different
polytypes of the low-pressure structure.

• Single-crystal samples maintain a high degree of texture both above the B1


phase transition and on back transformation, suggestive of a martensitic
transformation.

3
2) Boron Carbide

• Boron carbide was shocked compressed between 40 and 300 GPa. There is no
evidence for a crystallographic phase transition over this range.

• X-ray diffraction data provide evidence for partial amorphization which may
explain changes in the material’s continuum-level Hugoniot response.

• Our results indicate that B4C is more compressible than indicated by previous
lower pressure data.

• The effect of shock compression is to reduce the grain size and texture of the
sample producing more uniform, polycrystalline textures.

Introduction

B4C and SiC are high-strength materials of importance for the design of better lightweight
impact resistant coatings and tiles for satellites, spacecraft, space stations, vehicles and personal
armor [1–3]. These ceramics have outstanding mechanical properties, such as high hardness and
strength, and lower densities than metals, which make them ideal lightweight coating materials.
During ballistic impact conditions at pressures below 20 GPa, these ceramics exhibit high
strength, which allows them to form a rigid barrier and effectively break impacting projectiles.
However, at higher pressures, these materials are observed to have a strength collapse and
undergo possible phase transitions or amorphization. Currently, the mechanisms for the
catastrophic loss in strength are unknown as well as the lattice structure under compression. As a
result, both materials are problematic to simulate in the ballistic environment above 20 GPa.
Understanding their high-pressure response is a long sought after goal to improve the design and
performance of impact shielding [4]. Previous shock wave experiments have relied on interface
wave profiles that lacked the necessary in situ atomic lattice response information needed for
constitutive design models.

4
Figure 1. Shock Hugoniot (squares) and static (triangles) data of SiC [5]. Under shock
compression kinks in the Hugoniot at ~20 GPa and ~105 GPa have been observed. The 20-GPa
kink has been interpreted as the onset of plastic deformation. The 105-GPa discontinuity is
interpreted as a phase transition. Inset figure shows the calculated diffraction patterns for the
ambient cubic SiC structure and the high-pressure rocksalt phase [6] (as observed under static
compression).
Ceramic materials have many attractive properties including low density, high strength, high
melting point [7], low wear coefficient and high chemical stability that lead to their extensive use
in a wide range of industrial applications including as abrasives, shielding material in nuclear
reactors[ 8,9], high temperature electronic and thermoelectric devices[ 10], armor [11,12] and as
a potential ablator material for fusion capsules [13,14]. The effectiveness of high-strength
ceramics to resist high-pressure shock penetration has been an area of active research for many
years [15]. Silicon Carbide (SiC), for example, is used by NASA in the “Whipple” impact shield
design used to protect manned and unmanned spacecraft from collisions with micrometeoroids
and orbital debris whose velocities generally range between 3 and 18 km/s [16]. In this design,
two layers of SiC are separated by a high-strength carbon composite stand-off. The initial high-
strength SiC layer acts as a barrier that can shock an incoming particle and cause it to
disintegrate. This distributes the momentum over the inner SiC layer surface which acts as a
secondary barrier [3].
Silicon carbide is also important in geology and planetary science. It may be a host of
reduced carbon in the Earth’s interior as it is found in rocks from the mantle and in inclusions in

5
deep diamonds [17]. It also occurs in meteorites and impact sites. The first unambiguous
identification of SiC in meteorites was by Bernatowicz et al. [18] The unusual isotopic signature
[19–21] of SiC grains in meteorites indicate that they are pre-solar in origin and provide
constraints on stellar nucleosynthesis and on the stellar sources for the components of the solar
system. SiC is also a potential major constituent of some types of extrasolar planets. For
protoplanetary disks with high ratios of carbon to oxygen, the chemistry of planet formation may
lead to the formation of carbon-rich bodies. It has been reported that the mass and radius of the
exoplanet 55 Cancri e can be explained by a carbon-rich interior containing SiC [22]. Several
studies have explored possible interior structure of such carbon-rich planets [22,23], but

experimental data for the high-pressure phase of SiC are currently unavailable to test these
models.

Figure 2. Shock Hugoniot data for B4C [7]. The mechanisms underlying the observed kinks
at 20 and 40 GPa have been the subject of much controversy with suggestions of either a phase
transformation or amorphization. An additional kink at ~100 GPa has been observed in recent
Hugoniot experiments on the Omega laser facility by Fratanduono et al (2016) [24]. In contrast,
under static compression the ambient rhombohedral phase (inset) is observed to remain stable to
126 GPa [25].
At ambient conditions, there are over 250 known polytypes of SiC. These structures are
comprised of tetrahedrally coordinated SiC units with variable stacking sequences. The most
widely known among them are the hexagonal phase (6H polytype) and the cubic phase (3C
polytype). The equations of state of these polytypes have been measured up to 95-105 GPa using
a diamond anvil cell [6,26], and SiC has also been calibrated for use as a static pressure standard
[27]. The phase stability of SiC under pressure has been investigated where it has been observed
that the cubic 3C polytype transforms to a high-pressure phase having a rocksalt type structure at
about 105 GPa with a remarkably large volume change of 20.3% [6]. The transition is reversible
with a large hysteresis with the reverse transition occurring around 35 GPa upon unloading [6].

6
There has also been much interest in understanding the mechanism of transformation from the
tetrahedrally coordinated low-pressure phase to the octahedrally coordinated B1 phase, as these
types of transitions are common in semiconductors at high pressures. On the basis of theoretical
calculations, several different possible intermediate structures have been proposed [28–30].
Under shock compression, SiC has a moderately large Hugoniot Elastic Limit (HEL) (~20 GPa)
and appears to exhibit hardening above the HEL [31]. A phase transition was inferred under
shock loading at 105 GPa with a similar volume change as observed under static loading,
suggesting a similar structural change occurs under both static and dynamic loading (Fig. 1) [5].
In addition to SiC, Boron Carbide (B4C) is widely used as a high strength low-density
constituent of personal armor [11,12]. B4C is the third hardest material known (behind diamond
and cubic boron nitride) and it is the hardest material that is produced in tonnage quantities. B4C
is a member of the icosahedral boron compounds which are a unique family of crystalline solids
[10]. The building block of B4C is a B12 icosahedron that are held together by electron deficient
B-B bond (δ bond) [32] and a triatomic linker unit that bonds six neighboring icosahedra[10].
B4C crystallizes in rhombohedral R−3m space group, with 45 atoms in the unit cell (Fig. 2 inset)
[25]. Experimental challenges have limited the ability to determine the exact locations of the C
atoms in the crystal structure, and their positions are still widely debated [10]. Recent powder
diffraction diamond anvil cell experiments have shown that the ambient, rhombohedral phase
remains stable to quasi-hydrostatic compression up to pressures of 126 GPa [25]. The low
compressibility of B4C (K0 = 306 GPa) [33] is attributed to the force transfer between the rigid
icosahedral structural units through the inter-icosahedral bonds. Under static compression, the
icosahedral units compress less than the bulk of the crystal, retaining their close-to-ideal
geometry[33].
In contrast, the dynamic compression behavior of B4C from 10-70 GPa is significantly more
complex [7,34–40]. Recently the presence of two phase transitions at pressures of 20-40 GPa and
above 40 GPa have been postulated and debated based upon Hugoniot data and wave profile
analysis [7,39,41]. Arguments are focused around kinks in the Hugoniot associated with
pressure-induced volume collapses (see Fig. 2). It is unclear if the loss of strength is mechanical
in nature (amorphization) or due to a phase transition [42]. Zhang et al. [7] noted that the
proposed phase transition in B4C is unique when compared to the wave profiles of other brittle
materials with known phase transitions. In particular, the wave profiles in B4C do not exhibit the
characteristic step-like features often observed in phase transitions in ceramic materials. This
could be attributed to the unique crystal structure of B4C with strong covalent bonding and the
phase transition may result from the destruction of the linear triatomic linker unit. Recent
experiments by Yan et al. [43] have shown in DAC experiments that structure destabilization
occurs with quasi-uniaxial loading resulting in a bending of the triatomic linker and eventual
collapse of the crystal structure, while under quasi-hydrostatic compression,
It has been proposed that a possible high-pressure phase or amorphization in B4C might
explain the measured decrease in the ballistic performance at high impact rates and pressures
[44]. For example, it has been found that the performance of B4C under high-pressure shock

7
compression has been below expectations, especially compared to less hard and denser SiC
[11,12]. The experiments reported here are the first direct in situ measurement under dynamic
compression of the structural behavior of B4C at pressures up to 150 GPa.

Experimental Approach

Sample Description and Characterization

Silicon Carbide

The structure of SiC is a derivative of the diamond structure formed from Si and C atoms
linked in an sp3 bonding network. SiC exhibits a large number of polytypes based on different
stacking sequences. In addition, it readily incorporates defects and impurities. The major
polytypes are listed in Table 1. The cubic form adopts the B3 or Zincblende structure and is
known as β-SiC. The 2H polytype is the B4 or wurtzite structure. The hexagonal polytypes are
collectively known as α-SiC.

Table 1. SiC polytypes


Polytype Stacking a (Å) c (Å) Space Group
3C (Zincblende, β-SiC, B3) ABC 4.3596 - F43m
2H (Wurtzite, B4) ABAB 3.073 5.048 P63mc
4H ABCB 3.073 10.053 P63mc
6H ABCACB 3.073 15.11 P63mc

The samples used here consisted of (0001) single crystals (polytype 4H) grown by metal
organic chemical vapor deposition (MTI Corp). In addition, we also used polycrystalline samples
in the cubic 3C form (Mitsui Corp.). Samples were characterized at ambient conditions by
X-ray diffraction and Raman spectroscopy and the results were consistent with literature values.
The polycrystalline sample appears to contain extensive stacking faults.

Boron Carbide

B4C samples were obtained commercially (Inasco Corp.). The starting materials were
examined by X-ray diffraction, Raman spectroscopy, and scanning electron microscopy. Lattice
parameters and Raman spectra were consistent with literature values but both showed the
presence of excess C in the form of graphite inclusions distributed in the sample.

Dynamic Compression Experiments

8
Dynamic compression experiments were carried out at the Linac Coherent Light Source
(LCLS) of the Stanford Linear Accelerator Center (SLAC). The LCLS is an x-ray free electron
laser (FEL) which provides unprecedented x-ray brightness in an intense, short-duration pulse. It
delivers monochromatic x-rays tunable up to 11 keV with ~100 femtosecond pulses producing
1012-1013 photons per pulse. The peak brightness is nearly 8 orders of magnitude greater than
storage ring sources. Our experiments were conducted at the Matter in Extreme Conditions
(MEC) beamline.
Our target packages are illustrated schematically in Figure 3. SiC and B4C samples were cut
and polished to a thickness of 40-60 µm. The polycrystals were 1.5 x 1.5 mm in lateral
dimensions while the single crystals were 1.5 x 2 mm. The samples were then glued to polyimide
(CH) ablators (~75 µm thick) with an approximately 1-µm thick epoxy layer. In some cases,
(100) LiF windows (100-µm thick) were attached to the rear surface of the samples. For SiC
samples, a 0.1 µm-thick Al coating was deposited on the ablation-side of the CH, CH/sample
interface and the sample/LiF interface. The samples were mounted on target plates with a
common orientation for the single crystals.
At the Matter in Extreme Conditions (MEC) beamline, the samples were dynamically
compressed using ~10-20 J pulses from a 527 nm Nd:Glass laser system focused to a diameter of
~100-250 micrometers (Figure 3). The two lasers were incident at 15 degrees from normal and
overlapped onto the sample surface to generate ablation-driven shock compression waves that
propagated through the sample. The laser pulses were 10-15 ns long and the pulse shape (quasi-
flat-top) had a modest upward ramp to maintain a steady shock in the sample as the ablation
plasma expands. Experiments were performed both with and without phase plates. The thickness
of the CH was chosen to anneal out high-frequency structure in the pressure drive due to
imprinting from phase plate intensity speckle.
A line-focused VISAR (velocity interferometer system for any reflector) was used to
monitor the free surface or interface velocity of the sample. The VISAR has a 300-µm field of
view and a 20-ns sweep. For interface velocity measurements, a 100-µm thick LiF window, with
a reflective 0.1-µm coating of Al was glued onto the B4C (or SiC) to give a measure of the
sample interface velocity history. Pressures were determined by simulating the measured wave
profiles with a 1-D radiation hydrodynamics code (Hyades) or by standard impedance matching
techniques. Pressure was controlled by tuning the laser spot size (200-250 µm) and energy (13-
30 J). The VISAR also provided information on X-ray timing and positioning as well as any non-
planarity of the load.
Samples were probed with 0.146 nm (8.5 keV) free electron laser (FEL) x-rays and
diffraction data recorded on CSPad detectors (Fig. 3). The X-rays (SASE mode) were quasi-
monochromatic (0.5% ΔE/E) and consist of 80-fs pulses containing ~1012 photons. The x-ray
spot size was ~50 µm. We used three large format and two small format CSPAD detectors
allowing us to cover approximately 15-100 degree two-theta range. The data were integrated
azimuthally to obtain one-dimensional x-ray diffraction profiles. CSPads were calibrated using
Si3N4, LaB6, and CeO2.

9
Time scans were performed using the same drive conditions and nominally identical targets.
That is, the same pulse shape, energies, and duration were applied to nominally identical targets.
The only change with that for each shot in a time scan, diffraction data were collected at different
probe times after the shock wave entered the sample. Probe times extended up to 80 ns after
release. For each material under study (B4C or SiC) we carried out time scans at different
pressure states associated with the expected onset of the structural phase transformation(s).
At early times, we probe the sample before the shock wave reaches the back surface and so
we sample both shock-compressed and uncompressed material. The resultant X-ray diffraction
pattern (recorded on in-vacuum CSpads, integrating pixel array detectors) simultaneously
recorded information on both the ambient and transformed crystal structure and sample texture.
Examination of the crystal structure and texture at different times after shock compression onset
allowed us to construct a picture of the transformation timescales into the new phase. As
expected, this time dependence is a sensitive function of the over-pressurization of the
equilibrium phase transformation pressure[45].

Figure 3. Experimental setup. For a given laser power (sample pressure) the LCLS x-rays
were diffracted from the 60-µm thick B4C (or SiC) sample at different times after the onset of
shock compression to diagnose the evolution of changes in crystal structure and microstructure
(grain size and texture).
We have developed procedures and software to analyze the diffraction and wave profiles
obtained in this project. X-ray diffraction data recorded on the pixel array detectors (CSPads)
was projected to a common reference plane and analyzed with existing software such as
DIOPTAS. Laser interferometry records were analyzed using standard codes. Wavecode
simulations using Hyades were also performed. A large number of experiments were carried out
due to the high repetition rate of LCLS (1 shot every 3-7 minutes) and so significant volumes of
data required processing.

10
Figure 4. Top: VISAR interferogram recorded at SiC-LiF interface. The VISAR field of view is
300 micrometers. Bottom: Interface velocity profile determined from the interferogram above.

We encountered some technical problems during our experimental run in February 2016,
especially with regard to the long-pulse laser. For a 250-µm spot, the planarity of the drive was
observed to be a function of the location of the beam on the phase plate. The “best” location
appeared to change over the course of the experiment. As a result, we chose to shoot without
phase plates for the latter part of our run. Another issue was that for the same input waveform,
the pulse shape from one laser had a more square shape whereas the other laser produced a more
ramp-shaped pulse. Despite these technical issues, we were able to obtain high-quality
experimental data on both materials under study.

Results

Silicon Carbide

For SiC, three main pressure scans were recorded at 100, 120, and 185 GPa together with
more limited data for a few other pressures. An example of VISAR data for a representative
silicon carbide shot is shown in Fig. 4. The plateau associated with the phase transformation is at
a SiC-LiF particle velocity of ~ 3.1 km/s. This is consistent with SiC pressure of ~100 GPa. The
VISAR also records shock transit time as we can observe both the shock entering and exiting the
silicon carbide sample.

11
Figure 5. Left: Time series of compression data for 3C-SiC (polycrystalline) at 185 GPa. X-ray
probe times after the shock enters SiC are listed on the right. B1 peaks are marked with asterisks
and B1 densities are listed. Right: Projected images of diffraction data from the CSPad detector
corresponding to those on the left. Two-theta is the diffraction angle and phi is the azimuthal
angle around the X-ray beam.

Figure 6. d-spacing as a function of probe time for SiC compressed at 185 GPa.

12
Upon compression at 185 GPa, peaks corresponding to the B1 phase appear within 1 ns
of the shock entering the sample (Fig. 5). For this experiment, the shock transit time through the
sample was approximately 3 ns. Upon release at later times, the diffraction peaks broaden
reflecting the wider stress distribution in the sample as a release wave propagates back from the
free surface. The peaks also move to higher d-spacing as the pressure unloads. The B3-B1
transition is reversed as the average pressure in the sample drops rapidly. By 8.6 ns, the sample
has almost completely released and only a small trace of the high-pressure phase remains. Upon
the reverse phase transition, the recovered B3 phase also exhibits texture, further suggesting an
orientation relationship between the low- and high-pressure phases. However, this texture is lost
at late times (~14 ns) and smooth diffractions rings are observed (Fig. 5). The released sample at
late times is predominantly in the 3C (cubic) form of the staring material. We observe no
evidence for any intermediate phases during the forward or reverse transformation processes.
Figure 6 summarizes the evolution of d-spacings as a function of probe time after the shock
enters the sample.

Figure 7. Polycrystalline SiC, 250 µm spot. Time series of x-ray diffraction patterns collected
for a peak stress of ~120 GPa. X-ray probe times after shock enters SiC are listed to the right. B1
peaks and compressed 3C peaks are marked with asterisks and lines respectively and densities
are listed.

The time scan for a sample of polycrystalline SiC at a pressure of 120 GPa is shown in
Figure 7. In this case, the shock transit time is about 5 ns. During the compression phase, we
observe a complicated pattern that includes peaks from the uncompressed material

13
(polycrystalline, 3C) ahead of the shock front, peaks of compressed low-pressure phase and
peaks of the high-pressure B1 phase. We therefore infer that the sample at this pressures is
within the mixed-phase region where the low- and high-pressure phases coexist. This is
consistent with observations from gas gun data at longer timescales. Upon release, the B1 phase
disappears and the peaks of the 3C phase shift to higher d-spacings as the pressure is release.
Both the high- and recovered low-pressure phases show considerable texture upon
transformation.
Figure 8 shows the behavior of the single-crystal 4H sample. The pressure for this time
series of shots was close to 120 GPa based on analysis of VISAR records (pressure range
between 110-125 GPa). In this case, the transformation to the high-pressure B1 structure occurs
prior to the first image (1.9 ns). The reverse transformation begins at around 9.8 ns. The sample
retains a high degree of crystallinity after transition to B1 phase and initially upon reversion back
to the low-pressure structure. At very late times (43.1 ns), the peaks are smooth and untextured,
indicating a randomly oriented polycrystals.

Figure 8. Left: Time series of integrated x-ray diffraction patterns collected for a peak stress of
~120 GPa using a ~200-µm laser spot. The sample was a single crystal (4H polytype). X-ray
probe times after shock enters the SiC are listed on the right. In this case, the sample has already
converted to the B1 phase at the first recorded time (1.8 ns). B1 peaks are marked with asterisks
and corresponding densities are listed. The reverse transformation begins at 9.8 ns. Right:
Projected 2D images from CSPad detector. Sample retains a high degree of crystallinity after
transition to B1 phase and through reversion back to the low-pressure structure until late times.

14
Figure 9. LCLS vs gas gun data for SiC. Our results (Red) are compared to earlier studies (grey)
[5,40]. Our high-pressure datum near 185 GPa is denser than expected from gas gun studies, yet
is consistent with the calculated zero Kelvin B1 equation of state (EOS). The densities of the two
constituents in the mixed phase regime lie on the extrapolated Hugoniot curve and the B1 EOS
and a mixture of the 3C and B1 phases can explain the continuum density measurement.

Figure 9 shows a comparison of our LCLS results to earlier gas gun studies. At 120 GPa, we
observe coexistence of the B3 and B1 phases. The measured X-ray densities are consistent with
continuum Hugoniot measurements (6H phase) and theoretical calculations (B1 phase). The
mixed-phase region appears to be consistent with an approximately equimolar mixture of the two
phases. At higher pressures, we obtain a density of the B1 phase in general agreement with
theoretical calculations. Thus, our data are consistent with earlier studies. Calculated shock
temperatures at the phase transitions (~100-110 GPa) are 1300-1500 K on the Hugoniot and
~1000 K upon release [35].
Upon release from the B1 phase, SiC samples recrystallize as a mixture of different
polytypes. This is illustrated in Figure 10. The figure shows data for single-crystal (4H) shocked
into the B1 phase at ~120 GPa and released. A Rietveld refinement of the late-time data shows
that the released sample can be described as being composed of 52% of the hexagonal 4H
polytype and 48% of the cubic 3C polytype.

15
Figure 10. Profile refinement of shock-released SiC (4H). The peak shock pressure was 120
GPa. At late time (43 ns after shock enters the sample), the diffraction pattern can be fit be an
approximately equal mixture of 4H and 3C polymorphs indicating the sample reverts to a
mixture of the two main polytypes upon unloading.

Boron Carbide

X-ray diffraction time scans of shock-compressed boron carbide were carried out at two
primary pressures (40 GPa, 80 GPa) with limited additional data recorded at around 150 GPa and
300 GPa. Examples of projected CSPad images from each pressure range are shown in Fig. 11.

Figure 11. Representative CSPad data for B4C at four different pressures. Run numbers are
indicated.

The diffraction data for the uncompressed material is very spotty as a result of the larger
grain size of these crystals. Upon compression, the diffraction data exhibits progressively
smoother rings as pressure increases (Fig. 11), consist with a reduction in grain size in response
to shock compression.

16
Representative integrated diffraction patterns at different pressures are shown in Figure 12.
The bottommost pattern shows pre-shot data. Note the presence of a graphite peak near 3.4 Å.
The second from the bottom pattern (magenta color) shows the pattern obtained from a sample
compressed to 45 GPa and then released back to ambient pressure (recorded 30 ns after the shock
enters the sample). The recovered pattern matches the expected pattern of B4C, indicating that
the starting material is completely recovered. However, the graphite peak is not observed any
more, which may indicate that the excess C has dissolved into the B4C sample under shock
loading. Under compression the peaks shift to lower d-spacing (Fig. 12) as expected. The peaks
also broaden considerably which reflects a strain gradient within the sample. At 80 GPa and
above, there is evidence for a diffuse scattering feature at ~2.0-2.5 Å that may indicate partial
amorphization of the sample.

'I(R140_14ns) 80 GPa'
'I(R244_9ns) 150 GPa'
'I(R150_30ns) Release from 45 GPa'
300GPa (R373-8ns) 'I(R150PreShot)'
'I(R146_18ns)'#1
'I_R373(cspad1)'
6000

150GPa (R244-9ns)
Intensity (arb. units)
Intensity [AU]

80GPa (R140-14ns)

4000
(012)
(01-4)

45GPa (R146-18ns)
(021)

(003)
(110)

(01-1)
(02-5)

(02-2)
(12-1)

(113)
(024)

2000

Relesed (from 45GPa, R150-30ns)


Gr
R150-Preshot

2.0 2.5 3.0 3.5 4.0 4.5 5.0


d [A]

Figure 12. Representative integrated X-ray diffraction patterns for boron carbide. (hkl) values
for selected peaks are indicated.

Figure 13 compares pressure-density data for B4C from our study with previous shock
wave experiments using continuum techniques. Our results are consistent with the Hugoniot
curve of Fratanduono [24]. Our results also indicate that B4C remains largely crystalline to 300
GPa, possibly accompanied by amorphization of a portion of the sample. There is no evidence
for a crystalline phase transition up to the peak pressure of this study.

17
B 4C
350
B 4C
350
Hugoniot fit (Zhang 2006)
300 data1
Hugoniot fit (Zhang 2006)
Vogler (1984)
B 4C
300
350 data1
Hugoniot fit
250 Vogler (1984)
u
Hugoniot fit
250 Zhang (2006)
300 u
Pavlovski (1971) B 4C
Zhang (2006)
P H [GPa]

200
350 XRD - Present study
Pavlovski (1971)
Hugoniot fit (Fratnduono 2016)
250
P H [GPa]

200 XRD - Present study


Hugoniot
Hugoniotfitfit(Zhang 2006) 2016)
(Fratnduono
[GPa]

300
150 data1
PPHH(GPa)

150 200 Vogler (1984)


Hugoniot fit
250
100 u B4C
150 Zhang (2006)
100
Pavlovski (1971) 40

PH [GPa]
P H [GPa]

200
50 XRD - Present study
100 Hugoniot fit (Fratnduono 2016)
30
50 20

1500 50 10
2.5 3 3.5 4 4.5 5
0 2.6 2.7 2.8
2.5 3 3.5 [gr/cc] 4 4.5 [gr/cc] 5
100 0
2.5 3 [gr/cc]
3.5 4 4.5 5
50 Density (g/cm3)
[gr/cc]
Figure 13. Pressure-density Hugoniot relationships for B4C comparing results from earlier
continuum
0
2.5
measurements
3
to the X-ray
3.5
diffraction4data of this study.
4.5 5
[gr/cc]

Summary

In this study, we have successfully demonstrated the capability to use monochromatic x-rays
from a free electron laser to probe the crystal structure of shock compressed materials on
nanosecond timescales. This provides a powerful new capability to determine the lattice level
structure of materials under dynamic loading. The crystal structure is the most fundamental
property of a material and so its determination is of first-order importance for structural and
physical property measurements under extreme conditions. The ability to determine such
structures under in situ conditions is a major advance in shock compression science.
We have demonstrated that SiC transforms directly to the octahedrally coordinated B1
structure on the nanosecond timescale of laser-driven shock experiments and reverts to the
tetrahedrally coordinated ambient phase within nanoseconds of release. The data collected at 120
GPa exhibit diffraction peaks from both compressed low-pressure material and transformed B1
phase, consistent with the expected properties of a mixed-phase region. The data at 185 GPa
show a complete transformation to the B1 phase. Densities determined from diffraction peaks are
in agreement with an extrapolation of previous continuum data as well as theoretical predictions.
We find no evidence for any intermediate phases. It is also notable that a high degree of

18
crystallographic texture was retained in both the high-pressure phase as well as on back
transformation to the fourfold coordinated phase. This is observed in both single crystal and
polycrystalline samples. This indicates that there are strong orientational relationships between
the low- and high-pressure phases that can be interpreted to provide information about
transformation pathways between tetrahedral and octahedral coordination structures. These are
the subject of on-going investigation.
In the case of B4C, there is no evidence for a crystalline phase transition up to 300 GPa but
we do find evidence for possible partial amorphization of the sample. Upon unloading, we
observe a diffraction pattern consistent with boron carbide at ambient conditions, although
changes in sample texturing occur upon unloading reflect a loss of texture and a reduction in
grain size. The peak arising from graphite inclusions is not observed after shock loading,
suggesting that this carbon was incorporated into the structure of boron carbide. Our results are
consistent with recent continuum measurements and suggest that partial amorphization is likely
responsible for the anomalous behavior of B4C under shock compression.

Human Resources

This project provided training for two postdoctoral fellows at Princeton University: Dr. Sally
June Tracy and Dr. June Wicks (now Asst. Professor, Johns Hopkins University). The project
also supported an international collaboration with Dr. Binyamin Glam (Soreq, NRC, Israel).

Collaboration

This work was performed in collaboration with R. F. Smith, D. E. Fratanduono, A.


Fernandez Pañella, J. H. Eggert (all at Lawrence Livermore National Laboratory), C. A. Bolme
(Los Alamos National Laboratory), A. Gleason, H. J. Lee, A. MacKinnon, and F. Tavella (all at
Stanford Linear Accelerator Center), S. Speziale (GFZ Potsdam), K. Appel (XFEL), V. B.
Prakapenka (Chicago), and B. Glam (Soreq, NRC).

Publications and Presentations

Dufy, T. S., Ultrahigh pressure dynamic compression, American Geophysical Union Fall
Meeting, New Orleans, LA, Dec. 2017. (Invited)

Tracy, S. J., R. F. Smith, J. K. Wicks, D. E. Fratanduono, A. E. Gleason, C. A. Bolme, S.


Speziale, K. Appel, V. B. Prakapenka, A. Fernandez Pañella, H. J. Lee, A. MacKinnon, F.
Tavella, J. Eggert and T. S. Duffy, High-pressure phase transition in silicon carbide under shock
loading using ultrafast x-ray diffraction, American Geophysical Union Fall Meeting, New
Orleans, LA, Dec. 2017.

19
Duffy, T. S., Shock compression of geological materials Impacts, planet formation, and
phase transitions, Dynamic Compression Summer School, Argonne National Laboratory, August
2017. (Invited).

Tracy, S. J., R. F. Smith, J. K. Wicks, D. E. Fratanduono, A. E. Gleason, C. A. Bolme, S.


Speziale, K. Appel, V. B. Prakapenka, A. Fernandez Pañella, H. J. Lee, A. MacKinnon, F.
Tavella, J. Eggert and T. S. Duffy, High-pressure phase transition in silicon carbide under shock
loading using ultrafast x-ray diffraction, American Physical Society Topical Conference on
Shock Compression of Condensed Matter, St. Louis, MO, July 2017. (Invited)

Duffy, T. S. Structures of strong ceramics under dynamic compression, Fourth High-Power


Laser Workshop, Stanford Linear Accelerator Laboratory, Palo Alto, CA, October 2016 (Invited)

Duffy, T. S., Dynamic compression: From meteorites to exoplanets, HPSTAR, Shanghai


China, August 2016 (Invited).

References

[1] Levinshtein M E, Rumyantsev S L and Shur M S 2001 Properties of Advanced


Semiconductor Materials: GaN, AIN, InN, BN, SiC, SiGe (John Wiley & Sons)

[2] Pechenik A and Kalia R K 1999 Computer-Aided Design of High-Temperature Materials


(Oxford University Press on Demand)

[3] Christiansen E L and Center L B J S 2003 Meteoroid/debris shielding (National


Aeronautics and Space Administration, Lyndon B. Johnson Space Center)

[4] Branicio P S, Kalia R K, Nakano A, Vashishta P, Shimojo F and Rino J P 2008 Atomistic
damage mechanisms during hypervelocity projectile impact on AlN: A large-scale parallel
molecular dynamics simulation study Journal of the Mechanics and Physics of Solids 56
1955–1988

[5] Sekine T and Kobayashi T 1997 Shock compression of 6H polytype SiC to 160 GPa Phys.
Rev. B 55 8034–7

[6] Yoshida M, Onodera A, Ueno M, Takemura K and Shimomura O 1993 Pressure-induced


phase transition in SiC Phys. Rev. B 48 10587–90

[7] Zhang Y, Mashimo T, Uemura Y, Uchino M, Kodama M, Shibata K, Fukuoka K, Kikuchi


M, Kobayashi T and Sekine T 2006 Shock compression behaviors of boron carbide (B4C)
Journal of applied physics 100 113536–113700

20
[8] Ektarawong A, Simak S I, Hultman L, Birch J and Alling B 2014 First-principles study of
configurational disorder in B 4 C using a superatom-special quasirandom structure method
Physical Review B 90 024204

[9] Carrard M, Emin D and Zuppiroli L 1995 Defect clustering and self-healing of electron-
irradiated boron-rich solids Phys. Rev. B 51 11270–4

[10] Domnich V, Reynaud S, Haber R A and Chhowalla M 2011 Boron carbide: structure,
properties, and stability under stress Journal of the American Ceramic Society 94 3605–
3628

[11] Roberson, C and Hazell, P J 2003 Ceramic Armor and Armor Systems Ceramic
Transactions vol 51 pp 152–63

[12] Karandikar P G, Evans G, Wong S, Aghajanian M K and Sennett M 2009 A review of


ceramics for armor applications Adv Ceram Armor IV Ceram Eng Sci Proc 29 163–75

[13] Burnham A K, Alford C S, Makowiecki D M, Dittrich T R, Wallace R J, Honea E C, King


C M and Steinman D 1997 Evaluation of B 4 C as an Ablator Material for NIF Capsules
Fusion Science and Technology 31 456–462

[14] Mattsson C, Randall G C, Xu H, Streckert H, Hill C and Nikroo A 2013 Boron Carbide
Materials for Inertial Confinement Fusion APS Meeting Abstracts

[15] Grady D E 1998 Shock-wave compression of brittle solids Mechanics of Materials 29 181–
203

[16] Whipple, F. L. 1947 Meteorites and space travel. The astronomical journal 52 131

[17] Mathez E, Fogel R, Hutcheon I and Marshintsev V 1995 Carbon Isotopic Composition and
Origin of Sic from Kimberlites of Yakutia, Russia Geochim. Cosmochim. Acta 59 781–91

[18] Bernatowicz T, Fraundorf G, Ming T, Anders E, Wopenka B, Zinner E and Fraundorf P


1987 Evidence for interstellar SiC in the Murray carbonaceous meteorite Nature 330 728–
730

[19] Virag A, Wopenka B, Amari S, Zinner E, Anders E and Lewis R S 1992 Isotopic, optical,
and trace element properties of large single SiC grains from the Murchison meteorite
Geochimica et cosmochimica acta 56 1715–1733

[20] Hoppe P, Amari S, Zinner E, Ireland T and Lewis R S 1994 Carbon, nitrogen, magnesium,
silicon, and titanium isotopic compositions of single interstellar silicon carbide grains from
the Murchison carbonaceous chondrite The Astrophysical Journal 430 870–890

[21] Amari S, Hoppe P, Zinner E and Lewis R S 1992 Interstellar SiC with unusual isotopic
compositions-Grains from a supernova? The Astrophysical Journal 394 L43–L46

21
[22] Madhusudhan N, Lee K K M and Mousis O 2012 A possible carbon-rich interior in super-
earth 55 Cancri e Astrophys. J. Lett. 759 L40

[23] Wilson H F and Militzer B 2014 Interior Phase Transformations and Mass-Radius
Relationships of Silicon-Carbon Planets ApJ 793 34

[24] Fratanduono D E, Celliers P M, Braun D G, Sterne P A, Hamel S, Shamp A, Zurek E, Wu


K J, Lazicki A E, Millot M and Collins G W 2016 Equation of state, adiabatic sound speed,
and Gruneisen coefficient of boron carbide along the principal Hugoniot to 700 GPa Phys.
Rev. B 94 184107

[25] Fujii T, Mori Y, Hyodo H and Kimura K 2010 X-ray diffraction study of B4C under high
pressure Journal of Physics: Conference Series vol 215 (IOP Publishing) p 012011

[26] Bassett W A, Weathers M S, Wu T-C and Holmquist T 1994 Equation of state of SiC up to
68.4 GPa High-pressure science and technology—1993 vol 309 (AIP Publishing) pp 145–
148

[27] Zhuravlev K K, Goncharov A F, Tkachev S N, Dera P and Prakapenka V B 2013


Vibrational, elastic, and structural properties of cubic silicon carbide under pressure up to
75GPa: Implication for a primary pressure scale J. Appl. Phys. 113 113503

[28] Catti M 2001 Orthorhombic Intermediate State in the Zinc Blende to Rocksalt
Transformation Path of SiC at High Pressure Phys. Rev. Lett. 87 035504

[29] Miao M S and Lambrecht W R L 2003 Unified path for high-pressure transitions of SiC
polytypes to the rocksalt structure Phys. Rev. B 68 092103

[30] Karch K, Bechstedt F, Pavone P and Strauch D 1996 Pressure-dependent properties of SiC
polytypes Phys. Rev. B 53 13400–13

[31] Vogler T J, Reinhart W D, Chhabildas L C and Dandekar D P 2006 Hugoniot and strength
behavior of silicon carbide Journal of Applied Physics 99 023512

[32] Emin D 2006 Unusual properties of icosahedral boron-rich solids Journal of Solid State
Chemistry 179 2791–2798

[33] Dera P, Manghnani M H, Hushur A, Hu Y and Tkachev S 2014 New insights into the
enigma of boron carbide inverse molecular behavior Journal of Solid State Chemistry 215
85–93

[34] Wilkins M L and others 1968 Third progress report of light armor program UCRL-50460,
Lawrence Livermore Laboratory, Livermore, CA

[35] McQueen R G, Marsh J W, Taylor J W, Fritz J N and Carter W J 1970 The equation of
state of solids from shock wave studies High Velocity Impact Phenomena ed R Kinslow
(New York: Academic Press) pp 515–68

22
[36] Gust W H and Royce E B 1971 Dynamic yield strengths of B4C, BeO, and Al2O3
ceramics Journal of Applied Physics 42 276–295

[37] Grady D E 1994 Shock-wave strength properties of boron carbide and silicon carbide. Le
Journal de Physique IV 4 C8–385

[38] Holmquist T J, Templeton D W, Bishnoi K D and others 1999 A ceramic armor material
database (DTIC Document)

[39] Dandekar D P 2001 Shock response of boron carbide (DTIC Document)

[40] Vogler T J, Reinhart W D and Chhabildas L C 2004 Dynamic behavior of boron carbide
Journal of applied physics 95 4173–4183

[41] Grady D E 2002 Applied Research Associates report to US Army TACOM-TARDEC,


Warren, MI, for Contract No (DAAE07-01-P-L843 for project)

[42] Dandekar D, Ciexak J and Somayazulu M 2008 Compression and associated properties of
boron carbide Proceedings of the 26th Army Science Conference

[43] Yan X Q, Tang Z, Zhang L, Guo J J, Jin C Q, Zhang Y, Goto T, McCauley J W and Chen
M W 2009 Depressurization amorphization of single-crystal boron carbide Physical review
letters 102 075505

[44] Domnich V and Gogotsi Y 2003 Indentation-induced phase transformations in ceramics


High Pressure Surface Science and Engineering 443

[45] Smith R F, Eggert J H, Swift D C, Wang J, Duffy T S, Braun D G, Rudd R E, Reisman D


B, Davis J-P, Knudson M D and Collins G W 2013 Time-dependence of the alpha to
epsilon phase transformation in iron J. Appl. Phys. 114 223507

23

You might also like