Propellants Explo Pyrotec - 2022 - Li - Systematic Builder For All Atom Simulations of Plastically Bonded Explosives

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Research Article

doi.org/10.1002/prep.202200003 pep.wiley-vch.de

Systematic Builder for All-Atom Simulations of Plastically


Bonded Explosives
Chunyu Li+,[a] Brenden W. Hamilton+,[a] Tongtong Shen,[a] Lorena Alzate,[b] and Alejandro Strachan*[a]

Abstract: The shock to detonation transition in heteroge- on ideal and isolated defects. Hence, developed a method,
neous plastically bonded explosives is dominated by en- denoted PBXGen, to build realistic PBX microstructures for
ergy localization into hotspots that arise from the inter- all-atom simulations. PBXGen is generally applicable, and
action of the shockwave with microstructural features and we demonstrate it with two systems: an RDX-polystyrene
defects. The complex polycrystalline structure of these ma- PBX with a 3D microstructure and a TATB-polystyrene with
terials leads to a network of hotspot that can coalesce into columnar grains. The resulting structure exhibit key fea-
deflagration and detonation waves. Significant progress has tures of PBXs, albeit at smaller scales, and are validated
been made on the formation and potency of hotspots us- against experimental mechanical and shock properties.
ing atomistic simulations, but most of the work has focused
Keywords: Molecular dynamics simulation · Energetic materials · Plastically Bonded Explosives (PBX) · Microstructure · Hotspot formation

1 Introduction spots. In addition, these studies have shown that more nu-
merous, smaller pores lead to shorter run-to-detonations
The shock compression of materials can drive a variety of than fewer, large pores [18]. Despite this progress, the rela-
events such as phase transformations [1–3], plasticity [4–7], tive potency of the various mechanisms above are not
melting [8, 9], intra-molecular deformations [10–12], and known, nor are the possible interactions between them. For
chemical reactions [13–16]. Most of these phenomena can example, localized plastic deformation or interfacial shear
be exacerbated by the localization of energy due to the in- can weaken shocks and weaken the effect of porosity col-
teraction of the shockwaves with the materials micro- lapse. At the same time, the combination of neighboring
structure and defects. A key example of this is the for- hotspots caused by different mechanisms can result in larg-
mation of hotspots in composite energetic materials, which er and more reactive hotspots.
accelerates chemical reactions and controls initiation and The spatial and temporal scales of hotspots make their
sensitivity [6, 17–21]. In the shock initiation of energetic ma- experimental characterization extremely challenging, but
terials, the bulk shock temperature is typically not enough progress is occurring. For example, recent experimental
to prompt chemistry on fast enough timescales to cause a work from Bassett et al. have measured the peak temper-
run-to-detonation. Hotspots allows for more prompt re- atures of hotspots formed via shock compression, measur-
actions in small regions that can grow and coalesce into a ing values that are sensitive to microstructure and in the
deflagration wave and potentially a detonation [19]. Plastic range 4000–7000 K [28, 29]. However, these techniques cur-
bonded explosive (PBX) microstructures are often plagued rently can only measure the peak temperature values and
with voids, both between grains and within them, numer- are unable to directly extract the mechanisms responsible
ous cracks, grain boundaries, and various other defects that for hotspot formation.
will lead to complex hotspots upon shock compression Given the experimental challenges, computational mod-
[22, 23]. Several mechanisms are known to localize energy eling of hotspot formation has attracted significant interest
as shock induced hotspots [24], including the collapse of
porosity, shear band formation, crack formation, friction,
jetting, and viscous flow heating [5, 24–26]. [a] C. Li,+ B. W. Hamilton,+ T. Shen, A. Strachan
The collapse of porosity was found to be one of the School of Materials Engineering and Birck Nanotechnology Center,
Purdue University, West Lafayette, Indiana, 47907 USA
dominate mechanisms for hotspot formation through shock *e-mail: strachan@purdue.edu
desensitization experiments in which initial, weak compac- [b] L. Alzate
tion waves remove a significant fraction of the porosity, Computational Science and Engineering Division, Oak Ridge Na-
rendering the sample non-detonable from subsequent tional Laboratory, Oak Ridge, TN 37830 USA
shocks [27]. Comparisons of inclusions (silica microbeads) [+] Equal contributions.
and pores (air bubbles) in gelled nitromethane have also Supporting information for this article is available on the WWW
shown the superiority of the latter in forming critical hot- under https://doi.org/10.1002/prep.202200003

Propellants Explos. Pyrotech. 2022, 47, e202200003 (1 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

and provided much of our current mechanistic under- MD simulations are performed with LAMMPS. All energetic
standing. Hydrodynamic calculations provided a first pic- materials and polymer binders are described with non-re-
ture of the complexity of pore collapse [30–33]. With in- active, valence force fields that describe interactions in
creasing computational power and due to its explicit terms of covalent, van der Waals and electrostatic energies.
simulation of chemical reactions and thermo-mechanical Specifically, polystyrene (PS) is modeled with the Dreiding
processes with atomic detail, reactive molecular dynamics potential [64], TATB is modeled with the potential from
(MD) has been a key player in understanding and predict- Bedrov et al. [65–67], and RDX is modeled with the Smith-
ing the shock-induced reactivity [34–39]. For example, MD Bharadwaj potential [68] with the modifications from Ref
simulations revealed the importance of jetting and re- [40]. The cross pairs for the van der Walls interactions span-
compression as a mechanism of energy localization during ning two different force fields are added following the
pore collapse [26] and the importance of pore size and Dreiding potential rules. The atomic charges for PS are ob-
shape [40]. However, the large majority of MD simulations tained using the Gasteiger method [69] as implemented in
of hotspot formation and criticality have been limited to Polymer Modeler [70]. Long range electrostatic interactions
simplified geometries, such as single crystals with cylin- are calculated by the particle-particle particle-mesh (PPPM)
drical and diamond shaped pores, due to computation limi- method, which is an accurate and computationally efficient
tations [40–45]. To bridge the gap between atomistic simu- method developed by Hockney and Eastwood [71, 72]. A
lations and experimental measurements, continuum and timestep of 1.0 fs is used for RDX based systems and a time-
mesoscale work have also been applied to model hotspot step of 0.2 fs for TATB based systems. All boundary con-
formation mechanisms such as pore collapse, crack prop- ditions periodic except in the case of free surfaces, which
agation, and friction [25, 46–50]. These simulations have utilize shrink-wrapped boundaries.
been used to characterize simple isolated defects or micro-
structural features [51–56] and also realistic microstructures
[57–62]. However, these methods necessarily approximate 3 PBX Builder: PBXGen
many of the underlying mechanisms responsible for energy
localization, including localized plastic deformation, Typically, PBXs consist of 90–95 % HE particles, such as RDX,
amorphization, jetting, pore collapse, and recompression. HMX, or TATB, and 5–10 % polymer binder. Crystalline HE
Overall, numerous open questions remain regarding the particles are usually on the scale of micrometers to milli-
formation, interaction, and criticality of shock induced hot- meters, with a bimodal size distribution [23, 73]. Polymer
spots. binders with a relatively lower average molecular weight,
All atom simulations of realistic microstructures can be 2500–3000 g/mol, are preferred for processing [74]. The
the key to proving definite answers to these longstanding manufacturing process of a PBX has several key steps in-
issues of hotspot formation in shocked PBXs. While current cluding mixing HE in powder form with a polymeric resin,
computational capabilities restrict MD simulations to length curing the mixture at controlled temperature, and casting
scales below those in most PBX systems of interest, we be- into a desired shape under pressure [75].
lieve scaled down atomistic PBX models will provide in- PBXGen can build two classes of microstructures: 2D mi-
valuable insight to the relative importance of different hot- crostructures with columnar grains and 3D microstructures.
spot mechanisms. In this paper we introduce PBXGen, a While the latter is more realistic, the former enables larger
PBX microstructure builder for all-atom calculations, charac- feature sizes for a given computational budget. Regardless
terize the resulting microstructures for various composite of the microstructure type, the inputs include material se-
HE materials and use them with MD simulations to charac- lection (HE and binder types), relative amounts for each (by
terize their thermo-mechanical response. The main steps of weight percent), polymer chain molecular weight, HE crys-
the PBX builder are: coarse grain granular packing, HE grain tal shape and size distribution, and desired simulation cell
preparation, polymer coating of HE grains, and insertion of dimensions. With this information, the building process pro-
polymer-coated HE grains and system equilibration and ceeds along these steps: 1) Cutting HE particles from their
compaction. The resulting structures have densities and mi- single crystal structure with the desired shape and size dis-
crostructural features comparable to those of actual PBXs. tribution; 2) Coating HE particles with polymer chains fol-
Molecular dynamics simulations of the mechanical proper- lowing the desired length distribution and weight percent-
ties and hotspot formation in these structures reveal the age; 3) Packing coarse-grain particles representing coated
importance of inter-grain porosity with high aspect ratios to grains following the predetermined distribution; 4) Replac-
allow for significant molecular jetting. ing coarse-grain particles with all-atom coated particles fol-
lowed by a relaxation and compaction of the all-atom PBX
to the desired density under the desired thermodynamic
2 Simulation Methods conditions. Figure 1 illustrates these key steps to generate
an all-atom PBX system.
PBXGen is implemented by utilizing the LAMMPS [63] simu-
lation package for a majority of steps, and all subsequent

Propellants Explos. Pyrotech. 2022, 47, e202200003 (2 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

Figure 1. Flow chart of PBX builder: PBXGen.

3.1 Step 1: HE Crystalline Particles The first approach is based on surface energies and is a
modification to the Wulff construction to account for sto-
In this step, individual HE particles are created by cutting chasticity. Wulff found that minimizing surface energy leads
the desired geometries and sizes from periodic single crys- to the distance from the center of crystal to a given facet is
tals. The shape of HE particles used in PBX systems depend proportional to its surface energy and has been shown to
strongly on the processing method. The morphology result- generate faceting in good agreement with experiments in
ing from crystal growth has been extensively studied and the HE crystal HMX [79]. High energy surfaces that fall out-
several models have been proposed, see Ref. [76, 77] for re- side the convex hull do not appear and low-energy surfaces
cent reviews. Early work focused on minimizing the total have larger areas. In order to generate a family of particles
energy of a crystal for a give volume which leads to the with varying shapes we generate crystallites with a proba-
Wulff construction. Attempts to capture the kinetics of bility expðbEi Þ where b is an adjustable parameter that con-
growth include models based on attachment energy and trols fluctuationsPand Ei is the energy of crystal i and com-
their modifications, and more recent work builds on a puted as: Ei ¼ j Aj gj , where Aj is the area of facet j in
mechanistic understanding of growth. We note that HE par- crystal i and gj is the corresponding surface energy. The
ticles are often milled, which significantly affects their above distribution is generated using a Metropolis Monte
shape [78]. Our first effort on PBXGen implements two ap- Carlo algorithm [80]. Two moves are considered: one that
proaches in addition to simple spherical particles, one alters the radius of one of the free surfaces to the center of
based on surface energies and one on simplified kinetics via the grain and another that potentially alters which crystallo-
attachment energies. Other methods and software to create graphic free surface it is, choosing from a geometrically
crystalline particles can be easily integrated into PBXGen. neighboring surface in the set of available surfaces. The ra-

Propellants Explos. Pyrotech. 2022, 47, e202200003 (3 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

dius is altered using on a normal distribution and the sur- ticle, the number of polymer chains needed to give the cor-
face choice a uniform distribution, making both steps rever- rect weight percentage for the PBX is calculated. For the
sible. columnar microstructure, it is computationally easier to
The second approach implemented in PBXGen approx- generate polymer chains directly in the vacuum surround-
imates the crystal growth speed between the various surfa- ing the HE particle by using simple Configurational Bias
ces as proportional to the attachment energy [81]. In our Monte Carlo approach [83].
case, we define the attachment energy as the energy For the 3D microstructure, where significantly more pol-
evolved when a single molecule is added to a flat surface. ymer is needed, we start by placing coarse-grained (CG)
The more negative the attachment energy, the faster the beads, each representing a globular polymer chain, on the
corresponding surface will grow and the smaller area they surface of a larger CG particle representing the HE grain.
will take in the final structure. The attachment energy can The polymer CG beads are placed on the CG particle sur-
be computed using molecular mechanics and radii connect- face one by one, randomly, and avoiding overlap, until the
ing the center of the grain to surface i is proportional to desired number of CG beads is reached. If the pre-
1 Eatt determined amount could not be reached after a preset
Ri ¼ 2 D max i Eatt , where Eatt
i is the attachment energy of sur-
ði Þ maximum number of iterations, the radius of the CG par-
face i and D is the diameter of particle. Given selected par- ticle for the HE grain is slightly increased. After all CG beads
ticle diameter from the size distribution, the radii of each are placed, pre-equilibrated all-atom polymer chains are in-
face are then set to be proportional to the relative mor- serted into the CG bead locations and the all-atom HE grain
phology importance. is placed into the center of the large CG particle.
Figure 2 shows examples of grain morphologies created To accelerate the coating process, inwards velocities are
by PBXGen, with the rightmost stemming from the MC ap- assigned to all polymer atoms with velocities in the direc-
proach. The facetted RDX particle was obtained using the tion of the center of HE particle and the initial value in the
attachment energy approach and the columnar TATB crys- range of 0.5 km/s to 1.0 km/s. Converging velocities can be
tal was obtained with the Monte Carlo approach. All param- assigned multiple times with adiabatic (NVE) dynamics run
eters, surface and attachment energies, are included in the in between. The coated particle is then equilibrated under
supplementary material. The surface energies obtained NVE conditions, with rigid HE atoms and reassigning poly-
from TATB are in good agreement with available experi- mer velocities according to a Maxwell-Boltzmann dis-
ments [82]. The orientation of the columnar TATB grains tribution. The system is then fully thermalized with fully
was chosen to maximize anisotropy, it was desired to max- flexible atoms in both materials with dynamics run under
imize the anisotropy in the plane of the shock, so the po- isothermal-isochoric (NVT) conditions. Figure 3 shows coat-
tential surfaces are of the hkl style {N,0,M} where M and N ed, all-atom HE particles. We note that the CG bead proce-
are integers; these grains have perfect [010] texture. dure works only for short chain lengths were a spherical CG
bad can represent it. For higher molecular weight systems
in which entanglements and more complex chain geo-
3.2 Step 2: Coating HE Particles with Polymers metries exist, the former method of placing chains farther
away in the vacuum is required.
Experimentally, HE and polymer binders are typically mixed
via a slurry process in which polymer powders and HE pow-
ders mixed in an organic solvent [75]. However, on MD
timescales, the process of polymers wetting the surface of
HE grains must be artificially accelerated. Given a HE par-

Figure 2. Various HE particles cut from a set geometries (left two) and a Monte Carlo Approach (right).

Propellants Explos. Pyrotech. 2022, 47, e202200003 (4 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

Figure 3. Polymer-coated HE particles. Left: coated cylindrical TATB grain with 7 % PS by weight. Right: coated spherical RDX particle with
10 % PS by weight.

3.3 Step 3: Packing CG Microstructure mensions. Packing fractions can readily reach around 0.65.
A bimodal particle size distribution used to create one of
Due to the high computational cost of packing and com- the RDX-based PBX systems and the resulting packed CG
pressing the totality of the all-atom grains, an initial pack- system are shown in Figure 4.
ing using CG particles is performed, where the particles fol- Note that most PBXs exhibit a bimodal size distribution
low the size distributions of the coated all-atom HE grains. [23, 73] with large grains in the order of 100 μm and small
The CG grains are packed into the system cell using the grains in the range 0.5 to 5 μm. These sizes are not achiev-
granular dynamics package [84] as implemented in able with atomistic simulations and the example in Figure 4
LAMMPS. CG particles with different diameters are ran- shows a microstructure with a scaled down bimodal dis-
domly poured into the box and the box dimensions can be tribution affordable for all-atom MD simulations.
adjusted to make sure the entirety of the desired grains are
placed into the box.
After all CG particles are placed in the simulation cell,
granular dynamics under NPT conditions is conducted to
get a high packing fraction with the desired box di-

Figure 4. A packed system with CGs representing coated HE particles which diameters in a bimodal distribution.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (5 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

3.4 Step 4: Packing, Equilibration, and Compaction In some circumstances, such as impacting for shock-
waves, a system with a free surface or non-periodic boun-
The next step is to place the coated particles into the CG dary is desirable. To compress a PBX system whilst leaving a
microstructure. Each packed CG particle is replaced with a surface free involves setting a simulation cell with artificially
corresponding coated HE particle, and a resided rotation extended boundaries in one direction, without applying
can be applied, typically either random or no rotation. Fig- pressure or explicit deformation in that dimension. Other
ure 5 displays the final packed all-atom systems after CG re- dimensions, with periodic boundaries, are still compressed
placing. It can be seen that there are considerable spaces in as described above. For the non-periodic dimension, an in-
between coated HE particles. The overall mass density is ward velocity is applied on top of thermal velocities for
typically in the range of 0.3–0.4 g/cm3; therefore, additional grains near the free surfaces. This inward velocity is chosen
compaction is necessary. in the range 100–200 m/s. We find this to be a good com-
Compaction is the most computationally expensive part promise between the ability to compact the system within
of the building process as it is the only step that runs on reasonable computation time while not being high enough
the entire all atom system, requiring several hundred pico- to cause damage or expansion along the direction of the
seconds of run time to perform properly. Initially, the sys- open boundary conditions. This non-periodic compaction
tem is thermalized to a high temperature (~ 600 K) under potentially leads to a complex microstructure where the
isothermal isochoric conditions (NVT), keeping the HE central region of the system reaches a higher density, while
atoms fixed for ~ 20 ps. Isobaric isothermal (NPT) are then the regions close to free surfaces leaves with a higher level
conducted at moderate temperature (~ 450 K) and moder- of inter-grain porosity, see Figure 6b. This will be taken ad-
ate pressure (~ 50 MPa) bring the system to a compacted vantage of in this work to inspect the shock response of dif-
density without storing much strain energy within the in- ferent microstructures in a single simulation.
dividual grains. To save computational cost, the relaxation
is separated into three stages. In the first stage, the long-
range electrostatic interactions, which are very time con- 4 Mechanical Properties
suming, are turned off until the density reaches a stable val-
ue. In the second stage, the long-range electrostatic inter- PBXGen is utilized here to generate a set of 3D micro-
action is turned on and NPT dynamics continue until structures consisting of 1,3,5-Trinitro-1,3,5-triazinane (RDX)
equilibrium is again reached. Finally, the NPT dynamics are and Polystyrene (PS) with two different RDX/PS weight ra-
set to 300 K and 1 atmosphere for a final equilibration to tios: 83/17 and 90/10, and a columnar microstructure with
ambient conditions. For the RDX/PS system, these three 2,4,6-triamino-1,3,5-trinitrobenzene (TATB) and PS with the
stages take roughly 300 ps, 200 ps and 50 ps, respectively, TATB/PS weight ratio 93/7.
reaching a density of about 1.5 g/cm3. The columnar TATB system involves grain diameters fol-
lowing a normal distribution with an average of 15 nm and
a standard deviation of 2 nm. PS chains are all 40 mono-
mers in length. A free surface is left in the shock direction,
which is about 350 nm in length, with a width of 60 nm. In
the central, well packed region, the density is 1.73 g/cm3,
where theoretical max density is 1.93 g/cm3 and ex-
perimental samples of PBX 9502 can have densities near
1.89 g.cm3 [85].
The diameter of RDX particles range from 5 nm to
22 nm and number of particles of each size is determined
based on the bimodal distribution shown in Figure 4. The
PS chains are all 40 monomer long (molecular weight
~ 4168 g/mol). Four different RDX/PS systems are created,
varying RDX particle shape and orientation, as well as poly-
mer percentage. The cell dimensions are approximately
60 nm × 20 nm × 100 nm with the total number of atoms
around 10 million. Table 1 summarizes the various RDX-
based PBXs and the resulting mechanical properties.
The final densities of these systems are all approx-
imately 1.5 g/cm3. The ideal, void-free density for RDX/PS
systems with a weight ratio 90/10 is 1.72 g/cm3. However,
the maximum density typically reached experimentally is
Figure 5. Packed systems with all-atom coated HE particles (Dark: about 1.58 g/cm3 [76]. The smaller density in our systems as
HE particles; yellow: PS chains). compared to experiments is expected due to the smaller

Propellants Explos. Pyrotech. 2022, 47, e202200003 (6 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

Figure 6. Compacted all-atom PBX systems: (a) RDX/PS 3D system (Dark: RDX, yellow: PS), (b) TATB/PS pseudo-2D system with free surfaces
(Light: TATB, red: PS).

Table 1. PBX systems and mechanical properties.

PBX composition Density Porosity Young’s Tensile yield Tensile Shear Shear yield Shear yield
(g/cm3) (vol%) modulus (GPa) stress (MPa) yield modulus stress (MPa) strain (%)
strain (GPa)
(%)
RDX/PS: 83/17, spherical, 1.5085 5.65 8.1 � 1.0 148 � 8 5.9 � 0.9 1.9 � 0.1 87 � 5 8.4 � 0.6
oriented RDX
RDX/PS: 90/10, spherical, 1.4992 7.14 8.8 � 0.9 178 � 8 6.6 � 1.0 2.1 � 0.1 92 � 6 7.7 � 0.7
oriented RDX
RDX/PS: 90/10, spherical, 1.4926 8.06 7.3 � 0.7 157 � 7 6.5 � 1.1 3.0 � 0.1 92 � 4 7.5 � 0.6
random RDX
RDX/PS: 90/10, 1.4909 8.07 7.4 � 0.8 160 � 5 6.6 � 1.0 2.6 � 0.1 92 � 6 6.9 � 0.5
faceted, random RDX

grain size. The porosity for these systems ranges from 5.5 % (RDX 94 %, Exon461 6 %) [76]. The open symbol curve in
to 8 %, larger than the experimental values, in the range of Figure 7 is experimental data for PBX9501 (HMX 95 %, Es-
2–3 % [67]. We note that increasing the polymer binder tane 2.5 %, BDNPA-F 2.5 %) under compression [86].
fraction reduces porosity as more space between HE grains The strong role of microstructure is clear in the tensile
can be filled. yield stress, which varies from 148 MPa to 178 MPa. Com-
To quantify the mechanical properties of our micro- paring the two microstructures with oriented spherical
structures, uniaxial and shear deformations were conducted grains, we observe high strength with increasing crystal
at 300 K with a strain rate of 2 × 109 s 1. Figures 7 and 8 load. We attribute this to the larger percentage of polymer
show the stress-strain curves for uniaxial and shear de- with a lower strength. Finally, a reduction in strength is ob-
formations, respectively. The resulting values of moduli, served in microstructures with random crystal orientations.
yield stresses, and yield strains are listed in Table 1. The While additional analysis is needed to fully understand this
moduli are obtained by fitting the stress-strain curves at result, we attribute the lower strength to anisotropic elas-
low strains which is taken to be the elastic region. The ticity resulting in more heterogeneous stress distributions
Young’s moduli range from 7.3 to 8.8 GPa, in excellent in the randomly oriented PBX. The resulting stress local-
agreement with the experimental data 8.6 GPa for PBX9407 ization can result localized deformation and yield.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (7 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

a barostat and thermostat designed around the Hugoniot


jump conditions. Shock pressures studied range from 1 to
30 GPa with 20 ps of runtime and a time step of 0.25 fs.
Quantities of interest are averaged over the last 10 ps. Fig-
ure 9 displays the shock velocity vs. particle velocity, Us-Up,
temperature vs. Up, and pressure-Up relationships. All PBXs

Figure 7. Tensile stress-strain curves of RDX/PS systems (ex-


perimental data is for PBX9501 compression taken from Ref. [86]).

Figure 8. Shear stress-strain curves of RDX/PS.

5 Shock Response

Understanding energy localization into hotspots during


shock loading is one of the main open science questions
motivating our development of PBXGen. In this section, we
discuss our initial results of the response of our PBX sys-
tems to shock loading. Section 5.1 uses a computationally
efficient simulation technique to characterize the Hugoniot
equations of state of our RDX-based PBX systems. Sections
5.2 and 5.3 focus on explicit shock propagation simulations
in RDX and TATB systems, respectively.

5.1 RDX/PS Hugoniot Equation of State

Hugoniot curves are generated using the constant stress,


uniaxial, Hugoniotstat method [87]. This technique allows Figure 9. Shock Hugoniot for RDX/PS PBXs: (a) Shock velocity; (b)
the system to evolve to a desired shocked state by coupling Shock temperature; (c) Shock pressure.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (8 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

studied exhibit similar Hugoniot curves, indicating in-


sensitivity to microstructural details. This is because the Hu-
goniot depends on the compressive response of the materi-
al.
Due to the polymer binder, as expected, the PBX sys-
tems have considerably lower shock velocities than that of
single crystal RDX (under the same Hugoniot conditions
with Smith-Bharadwaj potential [68]). These results are also
slightly higher than the experimental data for an RDX based
PBX9407 [88]. It follows suit from the Us-Up relations that
single crystal RDX has a much higher pressure response. In-
terestingly, the shock temperature for the PBXs is slightly
higher than that of the single crystal RDX due to the in-
clusion of porosity in the PBX systems.

5.2 RDX/PS Hotspot Formation

There are two ways to simulate the propagation of shock


waves in molecular simulations. The first, and most com-
mon method, is ballistic impact where an explicit collision
between a projectile and a target are simulated using adia-
batic MD. This requires samples with free surfaces and in-
volves a rigid and infinitively massive piston [89]. An alter-
native approach which is applicable in 3D periodic systems,
is a converging shock method, in which the outer simu-
lation cell boundaries in the shock direction are driven in-
wards at a velocity of 2Up , remapping atomic velocities as
atoms cross the moving boundary [90]. To avoid the crea-
tion of free surfaces, we used this latter technique to shock
our RDX PBX samples. Figure 10 shows a temperature map
Figure 10. Hotspot distribution in a RDX/PS system (shock in the
of one of the RDX/PS systems with 10 % PS and oriented vertical direction).
spherical RDX particles. The particle velocity for this simu-
lation is 2.5 km/s and the temperature range shown is in
the range of 300 K to 2500 K. Despite the system being re-
plicated once in the shock direction, such that each wave until it reaches the closely packed central region of our
transverses the same microstructure, the waves are passing sample (40–60 ps). At this point, a two wave (elastic-plastic)
microstructural defects (mainly porosity) from two different shock structure can be seen, with plastic particle velocities
directions, leading to hotspots that are not identical in size mostly independent of the grain despite high levels of ani-
or temperature. sotropy in TATB [91]. In the final two frames, when the
shock front re-enters a porous region, molecular jetting into
voids accelerates the wave locally, which leads to the bands
5.3 TATB-PS: Shock Loading of higher velocity (red) regions seen in the 80 ps frame.
Figure 12 shows 1D wave profiles of temperature and
The columnar TATB/PS system, built with a free surface and density for the same 1.5 km/s shock. As expected, the tem-
a graded microstructure, is impacted utilizing a reverse bal- perature profiles in the outer sections of the plot (regions
listic setup [89] with an explicit piston (momentum mirror) with higher porosity) are considerably higher than in the
at the bottom boundary. TATB/PS particle velocity was set center (packed region) due to the formation of hotspots.
at 1.5 km/s. Figure 11 shows molecularly averaged (mono- Figure 13 shows a composite map of all molecular/mon-
mer averaged for PS) maps of particle velocity at various omer temperatures for the columnar TATB PBX right after
times in the simulation. For early times when the wave is in shock loading. This is performed by taking the position and
the in the first porous region (10–30 ps), the wave front is internal (rotational-vibrational) temperature of every ME
highly disparate across the width of the sample, which is molecule and polymer monomer 5 ps after being shocked,
likely to do with the large amount of surface roughness on where shock time is defined as the first time the molecule/
the impact plane, seen in the leftmost panel of Figure 11. monomer has a velocity in the shock direction greater than
The wavefront does not reach a clean, nearly planar shape 1.0 km/s where Up for the shock is 1.5 km/s. This shows the

Propellants Explos. Pyrotech. 2022, 47, e202200003 (9 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

Figure 11. Molecular particle velocity maps for various times for the TATB + PS PBX for a shock of 1.5 km/s. White line across panel b frames
correspond to locations of the white dashed lines in the unshocked full system of panel a.

Figure 12. 1D temperature and density profiles for the 1.5 km/s shock in the TATB + PS PBX which each curve representing 10 ps intervals
from 10–80 ps.

clear and distinct increase in temperature in the porous re- Figure 14 shows distributions of the temperatures of
gions, marked as outside of the dashed red lines. These hot- TATB molecules in the two regions at 5 ps after shock. Not
spots originate entirely from inter-granular pore collapse as only is the tail of the porous region much longer, but the
no defects are built within the grains during the building most probable temperature is roughly 100 K higher in the
process. A comparison of this temperature map with the in- porous region. On the 5 ps timescale, thermal transport will
itial microstructure, Figure 6(a), indicates that elongated only play a minor role in this. Most likely, the shock speed/
pores result in the highest temperatures, this is consistent particle velocity in these grains is increased due to the ejec-
with prior MD simulations [11, 40]. ta into the pores re-shocking downstream, leading to high-

Propellants Explos. Pyrotech. 2022, 47, e202200003 (10 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

Figure 13. Composite dump for the TATB + PS PBX with each molecule/monomer rendered at its center of mass and colored by rotation-
vibration temperature for 5 ps after each molecule/monomer experiences the shock.

that microstructure has a strong effect on ultimate tensile


strength of the PBXs, and as expected, a weaker effect on
elastic properties. We also performed shock simulations on
both the TATB and RDX systems. The shock velocity vs. par-
ticle velocity relationship in the RDX system is in good
agreement with experiments. Both the TATB and RDX sys-
tems are shocked at high impact velocities to induce sig-
nificant hotspots in which the resulting thermal local-
izations are correlated to the initial microstructures, with
inter-granular porosity playing a big role in the develop-
ment of large hotspots.

Acknowledgements
Figure 14. Distribution of HE molecular temperatures in the TATB +
PS PBX for times at 5 ps after each molecule is shocked. This work was supported by the U.S. Office of Naval Research, Mul-
tidisciplinary University Research Initiatives (MURI) Program under
Contract No. N00014-16-1-2557, program managers Chad Stoltz
and Kenny Lipkowitz.
er local temperatures in the grains as well as in the hot-
spots.

Data Availability Statement


6 Conclusions
Data available on request from the authors.
We introduced PBXGen, a generally applicable all-atom
builder for PBX microstructures for molecular dynamics. Our
method starts from the desired size distribution of the HE References
grains and cuts from single crystals. These crystals are then
coated with polymer of the desired molecular weight, with [1] K. Kadau, T. C. Germann, P. S. Lomdahl, B. L. Holian, Atomistic
the amount of polymer determined by mass ratio. Given Simulations of Shock-Induced Transformations and Their Ori-
the size of the coated particles, we use granular simulations entation Dependence in Bcc Fe Single Crystals, Phys. Rev. B:
Condens. Matter Mater. Phys. 2005, 72, 1.
to pack grains at a coarse grain representation. The final
[2] S. Zhao, T. C. Germann, A. Strachan, Atomistic Simulations of
step is to insert the all-atom polymer coated HE grains into Shock-Induced Alloying Reactions in NiAl Nanolaminates, J.
their corresponding CG grain and system is equilibrated Chem. Phys. 2006, 125, 164707.
and compacted. We demonstrated PBXGen by building a [3] K. Kadau, T. C. Germann, P. S. Lomdahl, B. L. Holian, Micro-
columnar PBX system based on TATB and multiple 3D RDX scopic View of Structural Phase Transitions Induced by Shock
based microstructures with faceted and spherical particles. Waves, Science 2002, 296, 1681.
We characterized mechanical properties of the RDX systems [4] E. Jaramillo, T. D. Sewell, A. Strachan, Atomic-Level View of In-
via uniaxial deformation simulations and compared the re- elastic Deformation in a Shock Loaded Molecular Crystal, Phys.
sponse against experimental stress-strain curves. We find Rev. B 2007, 76, 064112.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (11 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

[5] M. J. Cawkwell, T. D. Sewell, L. Zheng, D. L. Thompson, Shock- Measured by Ultra-Small Angle X-Ray Scattering, Propellants
Induced Shear Bands in an Energetic Molecular Crystal: Appli- Explos. Pyrotech. 2006, 31, 466.
cation of Shock-Front Absorbing Boundary Conditions to Mo- [23] C. B. Skidmore, D. S. Phililips, P. M. Howe, J. T. Mang, J. A. Ro-
lecular Dynamics Simulations, Phys. Rev. B: Condens. Matter Ma- mero, The Evolution of Microstructural Changes in Pressed
ter. Phys. 2008, 78, 1. HMX Explosives, in Eleventh International Detonation Sympo-
[6] M. P. Kroonblawd, L. E. Fried, High Explosive Ignition through sium (Snowmass, Colorado, 1998).
Chemically Activated Nanoscale Shear Bands, Phys. Rev. Lett. [24] W. C. Davis, High Explosives The Interaction of Chemistry and
2020, 124, 206002. Mechanics, Los Alamos Sci. 1981, 2, 48.
[7] R. Ravelo, T. C. Germann, O. Guerrero, Q. An, B. L. Holian, [25] J. K. Dienes, Q. H. Zuo, J. D. Kershner, Impact Initiation of Ex-
Shock-Induced Plasticity in Tantalum Single Crystals: In- plosives and Propellants via Statistical Crack Mechanics, J.
teratomic Potentials and Large-Scale Molecular-Dynamics Sim- Mech. Phys. Solids 2006, 54, 1237.
ulations, Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 1. [26] B. L. Holian, T. C. Germann, J. B. Maillet, C. T. White, Atomistic
[8] V. I. Levitas, R. Ravelo, Virtual Melting as a New Mechanism of Mechanism for Hot Spot Initiation, Phys. Rev. Lett. 2002, 89, 1.
Stress Relaxation under High Strain Rate Loading, Proc. Natl. [27] A. W. Campbell, J. R. Travis, The Shock Desensitization of Pbx-
Acad. Sci. USA 2012, 109, 13204. 9404 and Composition B-3, Los Alamos Natl. Lab. Prepr. Arch.
[9] S. Zhao, T. C. Germann, A. Strachan, Melting and Alloying of Ni Quant. Biol. 1985.
Al Nanolaminates Induced by Shock Loading: A Molecular Dy- [28] W. P. Bassett, D. D. Dlott, Shock Initiation of Explosives: Tem-
namics Simulation Study, Phys. Rev. B: Condens. Matter Mater. perature Spikes and Growth Spurts, Appl. Phys. Lett. 2016, 109,
Phys. 2007, 76, 1. 091903.
[10] B. W. Hamilton, M. P. Kroonblawd, C. Li, A. Strachan, A Hot- [29] W. P. Bassett, B. P. Johnson, N. K. Neelakantan, K. S. Suslick,
spot’s Better Half: Non-Equilibrium Intra-Molecular Strain in D. D. Dlott, Shock Initiation of Explosives: High Temperature
Shock Physics, J. Phys. Chem. Lett. 2021, 12, 2756. Hot Spots Explained, Appl. Phys. Lett. 2017, 111, 061902.
[11] B. W. Hamilton, M. P. Kroonblawd, A. Strachan, The Potential [30] R. Menikoff, Pore Collapse and Hot Spots in HMX, AIP Conf.
Energy Hotspot: Effects from Impact Velocity, Defect Geome- Proc. 2004, 706, 393.
try, and Crystallographic Orientation, J. Phys. Chem. C. 2022, [31] L. E. Fried, L. Zepeda-Ruis, W. M. Howard, F. Najjar, J. E. Reaugh,
126, 3743. The Role of Viscosity in TATB Hot Spot Ignition, AIP Conf. Proc.
[12] B. W. Hamilton, A. Strachan, Many-Body Mechanochemistry: In- 2012, 1426, 299.
tra-Molecular Strain in Condensed Matter Chemistry, Chem- [32] F. M. Najjar, W. M. Howard, L. E. Fried, M. R. Manaa, A. Nichols,
Rxiv. 2022. G. Levesque, Computational Study of 3-D Hot-Spot Initiation in
[13] B. W. Hamilton, M. N. Sakano, C. Li, A. Strachan, Chemistry Un- Shocked Insensitive High-Explosive, AIP Conf. Proc. 2012, 1426,
der Shock Conditions, Annu. Rev. Mater. Res. 2021, 51, 101. 255.
[14] A. Strachan, A. C. T. van Duin, D. Chakraborty, S. Dasgupta, [33] N. Bourne, E. F. John, Shock-Induced Collapse and Lumines-
W. A. Goddard, Shock Waves in High-Energy Materials: The Ini- cence in Cavities. Philos. Trans. R. Soc. London. Ser. A Math.
tial Chemical Events in Nitramine RDX, Phys. Rev. Lett. 2003, 91, Phys. Eng. Sci. 1999, 357, 295.
7. [34] M. N. Sakano, A. Hamed, E. M. Kober, N. Grilli, B. W. Hamilton,
[15] D. D. Dlott, M. D. Fayer, Shocked Molecular Solids: Vibrational M. M. Islam, M. Koslowski, A. Strachan, Unsupervised Learning-
up Pumping, Defect Hot Spot Formation, and the Onset of Based Multiscale Model of Thermochemistry in 1,3,5-Trinitro-
Chemistry, J. Chem. Phys. 1990, 92, 3798. 1,3,5-Triazinane (RDX), J. Phys. Chem. A 2020, 124, 9141.
[16] M. R. Manaa, E. J. Reed, L. E. Fried, N. Goldman, Nitrogen-Rich [35] M. S. Powell, M. N. Sakano, M. J. Cawkwell, P. R. Bowlan, K. E.
Heterocycles as Reactivity Retardants in Shocked Insensitive Brown, C. A. Bolme, D. S. Moore, S. F. Son, A. Strachan, S. D.
Explosives, J. Am. Chem. Soc. 2009, 131, 5483. McGrane, Insight into the Chemistry of PETN under Shock
[17] M. A. Wood, M. J. Cherukara, E. M. Kober, A. Strachan, Ultrafast Compression through Ultrafast Broadband Mid-Infrared Ab-
Chemistry under Nonequilibrium Conditions and the Shock to sorption Spectroscopy, J. Phys. Chem. A 2020, 124, 7031.
Deflagration Transition at the Nanoscale, J. Phys. Chem. C [36] M. R. Manaa, E. J. Reed, L. E. Fried, G. Galli, F. Gygi, Early
2015, 119, 22008. Chemistry in Hot and Dense Nitromethane: Molecular Dynam-
[18] D. M. Dattelbaum, S. A. Sheffield, D. B. Stahl, A. M. Dattelbaum, ics Simulations, J. Chem. Phys. 2004, 120, 10146.
W. Trott, R. Engelke, Influence of Hot Spot Features on the Ini- [37] E. J. Reed, A. W. Rodriguez, M. R. Manaa, L. E. Fried, C. M. Tarv-
tiation Characteristics of Heterogeneous Nitromethane, Int. er, Ultrafast Detonation of Hydrazoic Acid (HN 3), Phys. Rev.
Detonation Symp. LA-UR-10-0, 2010. Lett. 2012, 109, 1.
[19] C. A. Handley, B. D. Lambourn, N. J. Whitworth, H. R. James, [38] B. W. Hamilton, B. A. Steele, M. N. Sakano, M. P. Kroonblawd,
W. J. Belfield, Understanding the Shock and Detonation Re- I. F. W. Kuo, A. Strachan, Predicted Reaction Mechanisms, Prod-
sponse of High Explosives at the Continuum and Meso Scales, uct Speciation, Kinetics, and Detonation Properties of the In-
Appl. Phys. 2018, 5, 011303. sensitive Explosive 2,6-Diamino-3,5-Dinitropyrazine-1-Oxide
[20] S. Davis Herring, T. C. Germann, N. Grønbech-Jensen, Effects of (LLM-105), J. Phys. Chem. A 2021, 125, 1766.
Void Size, Density, and Arrangement on Deflagration and Det- [39] B. W. Hamilton, M. P. Kroonblawd, M. M. Islam, A. Strachan,
onation Sensitivity of a Reactive Empirical Bond Order High Ex- Sensitivity of the Shock Initiation Threshold of 1,3,5-Triamino-
plosive, Phys. Rev. B 2010, 82, 214108. 2,4,6-Trinitrobenzene (TATB) to Nuclear Quantum Effects, J.
[21] M. Sakano, B. W. Hamilton, M. M. Islam, A. Strachan, Role of Phys. Chem. C 2019, 123, 21969.
Molecular Disorder on the Reactivity of RDX, J . Phys. Chem. [40] C. Li, B. W. Hamilton, A. Strachan, Hotspot Formation Due to
Chem. Phys. 2018, 122, 27032. Shock-Induced Pore Collapse in 1,3,5,7-Tetranitro-1,3,5,7-Tetra-
[22] T. M. Willey, T. Van Buuren, J. R. I. Lee, G. E. Overturf, J. H. Kin- zoctane (HMX): Role of Pore Shape and Shock Strength in Col-
ney, J. Handly, B. L. Weeks, J. Ilavsky, Changes in Pore Size Dis- lapse Mechanism and Temperature, J. Appl. Phys. 2020, 127,
tribution upon Thermal Cycling of TATB-Based Explosives 175902.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (12 of 14) © 2022 Wiley-VCH GmbH
Systematic Builder for All-Atom Simulations of Plastically Bonded Explosives

[41] C. A. Duarte, C. Li, B. W. Hamilton, A. Strachan, M. Koslowski, [60] C. Miller, D. Olsen, Y. Wei, M. Zhou, Three-Dimensional Micro-
Continuum and Molecular Dynamics Simulations of Pore Col- structure-Explicit and Void-Explicit Mesoscale Simulations of
lapse in Shocked β -Tetramethylene Tetranitramine ( β -HMX) Detonation of HMX at Millimeter Sample Size Scale, J. Appl.
Single Crystals, J. Appl. Phys. 2021, 129, 015904. Phys. 2020, 127, 125105.
[42] Y. Long, J. Chen, Theoretical Study of the Defect Evolution for [61] Y. Wei, C. Miller, D. Olsen, M. Zhou, Prediction of Probabilistic
Molecular Crystal under Shock Loading, J. Appl. Phys. 2019, Shock Initiation Thresholds of Energetic Materials Through
125, 065107. Evolution of Thermal-Mechanical Dissipation and Reactive
[43] P. Zhao, S. Lee, T. Sewell, H. S. Udaykumar, Tandem Molecular Heating, J. Appl. Mech. 2021, 88, 1.
Dynamics and Continuum Studies of Shock-Induced Pore Col- [62] M. E. Fortunato, J. Mattson, E. Decarlos, J. P. Larentzos, J. K.
lapse in TATB, Propellants Explos. Pyrotech. 2020, 45, 196. Brennan, Pre- and Post-Processing Tools to Create and Charac-
[44] R. M. Eason, T. D. Sewell, Molecular Dynamics Simulations of terize Particle-Based Composite Model Structures, Army Res.
the Collapse of a Cylindrical Pore in the Energetic Material α- Labs 2017, Report No. ARL-TR-8213.
RDX, Journal of Dynamic Behavior of Materials 2015, 1, 423. [63] S. Plimpton, Fast Parallel Algorithms for Short-Range Molecular
[45] M. P. Kroonblawd, B. W. Hamilton, A. Strachan, Fourier-like Dynamics, J. Comput. Phys. 1995, 117, 1.
Thermal Relaxation of Nanoscale Explosive Hot Spots, J. Phys. [64] S. L. Mayo, B. D. Olafson, W. A. Goddard, DREIDING: A Generic
Chem. C 2021, 125, 20570. Force Field for Molecular Simulations, J. Phys. Chem. 1990, 94,
[46] N. Grilli, C. A. Duarte, M. Koslowski, Dynamic Fracture and Hot- 8897.
Spot Modeling in Energetic Composites, J. Appl. Phys. 2018, [65] D. Bedrov, O. Borodin, G. D. Smith, T. D. Sewell, D. M. Dattel-
123, 065101. baum, L. L. Stevens, A Molecular Dynamics Simulation Study of
[47] N. Grilli, M. Koslowski, The Effect of Crystal Anisotropy and Crystalline 1,3,5-Triamino-2,4,6- Trinitrobenzene as a Function
Plastic Response on the Dynamic Fracture of Energetic Materi- of Pressure and Temperature, J. Chem. Phys. 2009, 131, 224703.
als, J. Appl. Phys. 2019, 126, 155101. [66] N. Mathew, T. D. Sewell, D. L. Thompson, Anisotropy in Sur-
[48] C. A. Duarte, N. Grilli, M. Koslowski, Effect of Initial Damage face-Initiated Melting of the Triclinic Molecular Crystal 1,3,5-
Variability on Hot-Spot Nucleation in Energetic Materials, J. Triamino-2,4,6-Trinitrobenzene: A Molecular Dynamics Study, J.
Appl. Phys. 2018, 124, 025104. Chem. Phys. 2015, 143, 094706.
[49] A. Dandekar, Z. A. Roberts, S. Paulson, W. Chen, S. F. Son, M. [67] M. P. Kroonblawd, T. D. Sewell, Anisotropic Relaxation of Ideal-
Koslowski, The Effect of the Particle Surface and Binder Proper- ized Hot Spots in Crystalline 1,3,5-Triamino-2,4,6-Trini-
ties on the Response of Polymer Bonded Explosives at Low Im- trobenzene (TATB), J. Phys. Chem. C 2016, 120, 17214.
pact Velocities, Comput. Mater. Sci. 2019, 166, 170. [68] G. D. Smith, R. K. Bharadwaj, Quantum Chemistry Based Force
[50] J. K. Dienes, Discussion of a Statistical Theory of Fragmentation Field for Simulations of HMX, J. Phys. Chem. B 1999, 103, 3570.
Processes, Mech. Mater. 1985, 4, 337. [69] J. Gasteiger, M. Marsili, Iterative Partial Equalization of Orbital
[51] R. A. Austin, N. R. Barton, J. E. Reaugh, L. E. Fried, Direct Numer- Electronegativity-a Rapid Access to Atomic Charges, Tetrahe-
ical Simulation of Shear Localization and Decomposition Re- dron 1980, 36, 3219.
actions in Shock-Loaded HMX Crystal, J. Appl. Phys. 2015, 117, [70] B. P. Haley, N. Wilson, C. Li, A. Arguelles, E. Jaramillo, A. Stra-
185902. chan, Atomistic Simulations of Amorphous Polymers in the
[52] H. K. Springer, S. Bastea, A. L. Nichols, C. M. Tarver, J. E. Reaugh, Cloud with PolymerModeler, ArXiv: 1503.03894 2015.
Modeling The Effects of Shock Pressure and Pore Morphology [71] R. W. Hockney, J. W. Eastwood, Computer Simulation Using Par-
on Hot Spot Mechanisms in HMX, Propellants Explos. Pyrotech. ticles (CRC Press, 1998).
2018, 43, 805. [72] E. L. Pollock, J. Glosli, Comments on P3 M, FMM, and the Ewald
[53] M. A. Wood, D. E. Kittell, C. D. Yarrington, A. P. Thompson, Mul- Method for Large Periodic Coulombic Systems, Comput. Phys.
tiscale Modeling of Shock Wave Localization in Porous En- Commun. 1996, 95, 93.
ergetic Material, Phys. Rev. B 2018, 97, 1. [73] C. B. Skidmore, D. S. Phillips, S. F. Son, B. W. Asay, Character-
[54] N. K. Rai, S. P. Koundinyan, O. Sen, I. V. Schweigert, B. F. Hen- ization of HMX Particles in PBX 9501, AIP Conf. Proc. 2008, 579,
son, H. S. Udaykumar, Evaluation of Reaction Kinetics Models 579.
for Meso-Scale Simulations of Hotspot Initiation and Growth in [74] R. G. Stacer, D. M. Husband, Molecular Structure of the Ideal
HMX, Combust. Flame 2020, 219, 225. Solid Propellant Binder., Propellants Explos. Pyrotech. 1991, 16,
[55] A. Kapahi, H. S. Udaykumar, Dynamics of Void Collapse in 167.
Shocked Energetic Materials: Physics of Void-Void Interactions, [75] B. Nouguez, B. Mahé, P. O. Vignaud, Cast PBX Related Tech-
Shock Waves 2013, 23, 537. nologies for Im Shells and Warheads, Sci. Technol. Energ. Mater.
[56] M. S. Sellers, M. Lísal, I. Schweigert, J. P. Larentzos, J. K. Brenn- 2009, 70, 135.
an, Shock Simulations of a Single-Site Coarse-Grain RDX Model [76] J. Li, C. J. Tilbury, S. H. Kim, M. F. Doherty, A Design Aid for
Using the Dissipative Particle Dynamics Method with Re- Crystal Growth Engineering, Prog. Mater. Sci. 2016, 82, 1.
activity, AIP Conf. Proc. 2017, 1793, 1. [77] Y. Zhao, C. J. Tilbury, S. Landis, Y. Sun, J. Li, P. Zhu, M. F. Doh-
[57] M. Akiki, T. P. Gallagher, S. Menon, Mechanistic Approach for erty, A New Software Framework for Implementing Crystal
Simulating Hot-Spot Formations and Detonation in Polymer- Growth Models to Materials of Any Crystallographic Complex-
Bonded Explosives, AIAA J. 2017, 55, 585. ity, Cryst. Growth Des. 2020, 20, 2885.
[58] Y. Wei, R. Ranjan, U. Roy, J. H. Shin, S. Menon, M. Zhou, In- [78] R. Ho, M. Naderi, J. Y. Y. Heng, D. R. Williams, F. Thielmann, P.
tegrated Lagrangian and Eulerian 3D Microstructure-Explicit Bouza, A. R. Keith, G. Thiele, D. J. Burnett, Effect of Milling on
Simulations for Predicting Macroscopic Probabilistic SDT Particle Shape and Surface Energy Heterogeneity of Needle-
Thresholds of Energetic Materials, Computational Mechanics Shaped Crystals, Pharm. Res. 2012, 29, 2806.
2019, 64, 547. [79] S. Yang, I. Bier, W. Wen, J. Zhan, S. Moayedpour, N. Marom,
[59] Y. Wei, S. Kim, Y. Horie, M. Zhou, Quantification of Probabilistic Ogre: A Python Package for Molecular Crystal Surface Gen-
Ignition Thresholds of Polymer-Bonded Explosives with Micro- eration with Applications to Surface Energy and Crystal Habit
structure Defects, J. Appl. Phys. 2018, 124, 165110. Prediction, J. Chem. Phys. 2020, 152, 244122.

Propellants Explos. Pyrotech. 2022, 47, e202200003 (13 of 14) © 2022 Wiley-VCH GmbH
Research Article C. Li, B. W. Hamilton, T. Shen, L. Alzate, A. Strachan

[80] N. Metropolis, S. Ulam, The Monte Carlo Method, J. Am. Stat. Rate and Temperature, in Eleventh International Detonation
Assoc. 1949, 44, 335. Symposium (No. LA-UR-98-3059; CONF-980803-) 1998.
[81] Y. Wang, X. Li, S. Chen, X. Ma, Z. Yu, S. Jin, L. Li, Y. Chen, Prepa- [87] R. Ravelo, B. L. Holian, T. C. Germann, P. S. Lomdahl, Constant-
ration and Characterization of Cyclotrimethylenetrinitramine Stress Hugoniotstat Method for Following the Dynamical Evo-
(RDX) with Reduced Sensitivity, Materials (Basel). 2017, 10, 974. lution of Shocked Matter, Phys. Rev. B: Condens. Matter Mater.
[82] J. D. Yeager, A. M. Dattelbaum, E. B. Orler, D. F. Bahr, D. M. Dat- Phys. 2004, 70, 1.
telbaum, Adhesive Properties of Some Fluoropolymer Binders [88] S. P. Marsh, LASL Shock Hugoniot Data (Vol. 5)., 1980.
with the Insensitive Explosive 1,3,5-Triamino-2,4,6-Trini- [89] B. L. Holian, G. K. Straub, Molecular Dynamics of Shock Waves
trobenzene (TATB), J. Colloid Interface Sci. 2010, 352, 535. in Three-Dimensional Solids: Transition from Nonsteady to
[83] J. Ilja Siepmann, D. Frenkel, Configurational Bias Monte Carlo: Steady Waves in Perfect Crystals and Implications for the Ran-
A New Sampling Scheme for Flexible Chains, Mol. Phys. 1992, kine-Hugoniot Conditions, Phys. Rev. Lett. 1979, 43, 1598.
75, 59. [90] B. L. Holian, D. E. Grady, The Microscopic “Big Bang”, Phys. Rev.
[84] N. V. Brilliantov, F. Spahn, J. M. Hertzsch, T. Pöschel, Model for Lett. 1988, 60, 1355.
Collisions in Granular Gases, Phys. Rev. E 1996, 53, 5382. [91] P. Zhao, M. P. Kroonblawd, N. Mathew, T. Sewell, Strongly Ani-
[85] R. L. Gustavsen, S. A. Sheffield, R. R. Alcon, Measurements of sotropic Thermomechanical Response to Shock Wave Loading
Shock Initiation in the Tri-Amino-Tri-Nitro-Benzene Based Ex- in Oriented Samples of the Triclinic Molecular Crystal 1,3,5-Tri-
plosive PBX 9502: Wave Forms from Embedded Gauges and amino-2,4,6-Trinitrobenzene, J. Phys. Chem. C 2021, 125, 22747.
Comparison of Four Different Material Lots, J. Appl. Phys. 2006,
99, 114907.
[86] G. T. Gray III, D. J. Idar, W. R. Blumenthal, C. M. Cady, P. D. Pe- Manuscript received: January 4, 2022
terson, High-and Low-Strain Rate Compression Properties of Revised manuscript received: March 17, 2022
Several Energetic Material Composites as a Function of Strain Version of record online: June 8, 2022

Propellants Explos. Pyrotech. 2022, 47, e202200003 (14 of 14) © 2022 Wiley-VCH GmbH

You might also like