You are on page 1of 11

Research Article

doi.org/10.1002/prep.202200001 pep.wiley-vch.de

Machine-Learning a Solution for Reactive Atomistic


Simulations of Energetic Materials
Rebecca K. Lindsey,*[a] Cong Huy Pham,[a] Nir Goldman,[a, b] Sorin Bastea,[a] and Laurence E. Fried[a]

Abstract: Many of the safety and performance-related body expansion of interatomic interactions. ChIMES has
properties of energetic materials (EM) are related to com- been successfully applied to a variety of systems including
plex condensed phase chemistry at extreme P,T conditions simple model energetic materials, both as a correction for
eluding direct experimental investigation. Atomistic simu- simpler quantum theory and as a stand-alone interatomic
lations can play a vital role in generating insight into EM potential. In this perspective, the successes and challenges
chemistry, but they rely critically on the availability of suit- of applying the ChIMES approach to the reactive molecular
able interatomic potentials (“force fields”). The ChIMES ma- dynamics of energetic materials are outlined. Our machine-
chine learning approach enables generation of interatomic learned approach is general and can be applied to a variety
potentials for condensed phase reacting systems, with ac- of different application areas where atomic-level calcu-
curacy similar to Kohn-Sham density functional theory lations can be used to help guide and elucidate experi-
through its unique, highly flexible orthogonal basis set of ments.
interaction functions and systematically improvable many-
Keywords: atomistic simulations · machine learning · reactive force field · interatomic model · ChIMES · chemistry

1 Introduction chanics (MM) simulations using interatomic potential (IAPs)


[10, 11] capable of modeling EM on spatiotemporal scales
Atomistic simulations have proven to be a critical tool for approaching 1 μm and 1 μs. Non-reactive MM IAPs, such as
energetic materials (EM) research over the past several dec- those developed by Bedrov et al. [12] are highly efficient
ades, since they can provide insight into processes occur- and accurate for non-reactive processes such as thermal
ring under the extreme and rapidly changing thermody- transport in solid energetics [13] while reactive MM IAPs
namic conditions associated with EM function, where direct (e. g., the widely used ReaxFF [14]) have been applied to a
experimental probing is often difficult or impossible [1]. variety of energetic molecules and processes [4, 10, 11].
This simulation approach has been applied to a wide varie- Although strides in the scientific communities’ under-
ty of application areas including predicting material equa- standing of EM evolution under extreme conditions have
tion of state (EOS) [2], determining nanoscale phenomena been made using these approaches, increased interest in
governing material evolution at the grain scale [3], and pro- high accuracy modeling of reactive processes in EM on larg-
viding insight into the mechanisms controlling EM safety, er spatiotemporal scales demands more of atomistic simu-
such as the formation of hot spots under shock loading [4– lation capabilities. For example, there is a need to rapidly
6]. Ultimately, atomistic simulations could accelerate efforts evaluate new and/or candidate materials via highly accu-
toward developing new energetic materials with improved rate and efficient crystal structure searches [15], understand
safety/energy trade-offs [7]. hot spot initiation mechanisms on experimental scales [16],
Nevertheless, atomistic simulations present their own and elucidate carbon condensation mechanisms in non-ide-
challenges. EM evolution under extreme conditions is large- al EM [17]. All of these applications require an agile model
ly governed by chemical reactivity occurring on the nano- development capability; however, standard MM IAP devel-
meter and picosecond spatiotemporal scales, yet character-
istic scales for relevant emergent phenomena (e. g., defect
formation, pore collapse, and carbon condensation) can [a] R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried
span up to four orders of magnitude [8]. Reconciling the in- Physical and Life Sciences Directorate
herently multiscale nature of EM research has traditionally Lawrence Livermore National Laboratory
required atomistic simulators to choose between highly Livermore, California 94550, USA
*e-mail: rklindsey@llnl.gov
predictive yet computationally intensive first principles (FP) [b] N. Goldman
approaches like Kohn-Sham [9] density functional theory Department of Chemical Engineering
(i. e., which confine accessible scales to a few hundred University of California, Davis
atoms and 10s of picoseconds), or efficient molecular me- Davis, California 95616, USA

Propellants Explos. Pyrotech. 2022, 47, e202200001 (1 of 11) © 2022 Wiley-VCH GmbH
Machine-Learning a Solution for Reactive Atomistic Simulations of Energetic Materials

opment approaches tend to be slow and yield models of ature to 1000 s of K and 10,000s of MPa, which can drive
either limited chemical transferability or accuracy. Ulti- the system from insulating to metallic [32], induce chemis-
mately, IAPs capable of bridging the gap between accuracy try, and result in phase separation [33, 34]. At the same
and efficiency of FP and MM methods are needed before time, energetic materials can include four or more ele-
these problems can be gainfully pursued. ments, which further complicates chemistry, necessitates
Machine learning (ML), or more specifically, machine- determining 103–105 of model parameters [35], and compli-
learned interatomic potentials (ML–IAPs) provide a promis- cates generation of a suitable training dataset. Ultimately,
ing path forward. Machine learning encompasses a broad developing ML–IAPs for EM is a difficult task. In this per-
class of data analytic approaches wherein the fitting frame- spective, practical challenges in EM ML–IAP development
work “learns” to construct models that make predictions are presented along with an overview of possible solutions
based on some set of data descriptors, with the minimal and outcomes, and discussion of remaining challenges and
human intervention [18]. Within the context of IAPs, the next steps, within the context of the ChIMES approach.
“data descriptors” generally summarize the atomic coor-
dinates, (e. g., pair distances, bending angles, and embed-
ding densities in the case of MM-IAPs), and are used to pre- 2 Methods
dict system energy, atomic forces, and system stress.
Quantum simulations can be used to generate the training 2.1 ChIMES Overview: Functional Form
data set, and parameters can be calibrated via, e. g., the
“force matching” approach [19]. Traditional MM–IAPs relate The goal in developing ChIMES is to establish a practical
a given set of atomic coordinates to the system potential and tractable alternative to traditional IAPs that enables:
energy through a fixed mathematical expression. Typically, * Efficient and accurate simulations of novel HE

this expression has a characteristic potential energy func- * Minimized time to determine optimal model parameters

tion shape and a fixed number of parameters. In contrast, * Maximized simulation efficiency

ML–IAPs use a collection of basis functions to represent the * Description of complicated reactive phenomena

potential energy and contain many more parameters (e. g., The ChIMES formalism draws upon insights from
typically in the range 103–105) than is typical for MM–IAPs, chemistry that bonding can be described in terms of a rela-
making them profoundly more flexible. For example, a sin- tively small number of interacting atoms. This suggests that
gle ML–IAP framework can be used to describe non-re- an expansion in the number of interacting atoms could
active, reactive, metallic, and/or insulating systems [20–23], yield a functional form that is simpler, faster to parameter-
whereas MM–IAPs generally require a different set of equa- ize, and more computationally efficient than common ANN
tions to describe each different system class. Note that this approaches.
flexibility is particularly helpful in cases where the appro- The design philosophy behind ChIMES involves map-
priate function form isn’t known a priori, which is partic- ping quantum mechanical energies onto linear combina-
ularly relevant for the extreme conditions produced by tions of many-body Chebyshev polynomials of the first
functioning energetic materials. ML–IAPs are generally ca- kind. These polynomials have a number of desirable proper-
pable of arbitrary complexity and high accuracy, although ties for creating interatomic potential energy surfaces, in-
electronic degrees of freedom such as spin are typically ne- cluding: (i) they are orthogonal with respect to a weighting
glected. In addition, ML–IAPs can be developed as correc- function and can be generated recursively, allowing for ba-
tions to approximate quantum methods such as DFTB [24– sis set completeness and user-defined complexity, (ii) high-
28] or to MM–IAPs such as ReaxFF [29]. Typically, correction er-order polynomials tend to have decreasing expansion
approaches require fewer parameters and simpler func- coefficient values (due to their monic form), and (iii) they
tional forms than ML–IAPs that completely describe the IAP. are “nearly optimal” in that the error in an expansion will
The advent of ML–IAPs has had a transformative effect closely resemble a minimax polynomial. In addition, de-
on broader atomistic simulation research and has stimu- rivatives of these polynomials are related to Chebyshev pol-
lated the development of diverse underlying forms and ynomials of the second kind, which themselves are orthog-
methods [20, 29, 30]. Artificial neural network (ANN) ap- onal, and can also be generated recursively. This allows for
proaches [31] have experienced a surge of popularity due easy and reliable determination of forces and stress tensor
to widely available open-source toolkits. However, the bulk components for atomistic calculations.
of model development and application efforts have focused Briefly, the ChIMES total energy corresponds to an n-
on inorganic/solid-state materials or materials under mild body expansion:
conditions relative to those associated with EM function.
EM ML–IAP development presents several unique chal-
lenges due to the need to describe condensed phase re-
activity in chemically complex organic systems. In partic-
ular, shocked materials experience a broad range of
thermodynamic conditions, varying from ambient temper-

Propellants Explos. Pyrotech. 2022, 47, e202200001 (2 of 11) © 2022 Wiley-VCH GmbH
Research Article R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried

X
na X
na X
na the f c smoothly varying cutoff functions for each con-
E nB ¼ 1
E i1 þ 2
E i1 i2 þ 3
E i1 i2 i3 þ � � � stituent pair distance. Penalty functions are not included in
i1 i1 >i2 i1 >i2 >i3 this case and instead are handled entirely by the two-body
X
na (1) interaction. Note that additional details on extending
nB
þ Ei1 i2 ���in B
ChIMES to greater than three-body interactions can be
i1 >i2 ��� inB 1 >inB found in [57].
Optimal ChIMES parameters (i. e., polynomial co-
efficients) are generated through nominal “force matching”
where EnB is the total ChIMES system energy, nB is the max- to per-atom force components, overall energies, and overall
imum bodiedness, n Ei1 i2 :::in is the n-body ChIMES energy for a stress tensors for configurations from DFT-MD simulations.
given set of atoms with indices i ¼ fi1 ; i2 ; :::; in g, and na is Since ChIMES models are parametrically linear, this is ac-
the total number of atoms in the system. The one-body en- complished by solving the overdetermined matrix equation
ergies, 1 Ei1 , correspond to the atomic energy constants for AC ¼ B, where the matrix A corresponds to the values of
each element type. the requisite polynomials for a given training configuration
The two-body (pairwise) energies are expressed as linear (i. e., derivatives with respect to the fitting coefficients). The
combinations of Chebyshev polynomials of the first kind: column vectors C and B correspond to the linear ChIMES
coefficients and the training forces, energies, and stresses,
� �X
O2B respectively. Note that this linear least-squares optimization
2 e e �
Ei1 i2 ¼ f p ri1 i2 þ f eci ei ri1 i2 Cemi ei T m si1i i2 i
1 2 1 2 1 2
(2) problem can be solved with any number of algorithms [28].
m¼1 The ChIMES cluster-based description thus provides a
chemically-motivated means of explicit many-body descrip-
e e �
In this case, T m si1i i2 i represents a Chebyshev poly- 1 2
tion while the Chebyshev polynomials afford a flexible, con-
e e
nomial of order m, and si1i i2 i is the pair distance transformed 1 2
vergent orthogonal basis with several mathematically favor-
to exist over the interval ½ 1; 1� using � a Morse-like function able features. Moreover, ChIMES models are parametrically
e e
[36]. Here, si1i i2 i / exp ri1 i2 =le1 e2 and le1 e2 is an element-
1 2
linear, hence advanced linear solvers (e. g., which perform
pair distance scaling constant, usually taken to be the peak parameter selection) can be used to rapidly generate mod-
position of the first coordination shell. Ceni ei is the corre- 1 2
els. In addition to use as a stand-alone ML–IAP, note that
sponding permutationally invariant coefficient for the inter- ChIMES can be used to generate corrections to other IAPs,
action between atom types ei1 and ei2 taken from�the set of e. g., as a Density Functional Tight Binding correction en-
all possible element types, feg. The term f eci ei ri1 i2 is a Ters- 1 2
abling application to materials under extreme conditions.
off cutoff function [37] which forces 2 Ei1 i2 to zero beyond a
maximum distance defined for a given fe1 ; e2 g pair type. To
prevent sampling of ri1 i2 distances below � the model inner 2.2 ChIMES Overview: Use as a DFTB Correction
cutoff, a smooth penalty function f p ri1 i2 is introduced. The
reader is referred to previous work for additional details The Density Functional Tight Binding method (DFTB) [38] is
[22]. a semi-empirical quantum simulation approach that can be
Greater than two-body orthogonal basis sets are gen- viewed as an efficient proxy for full-quantum calculations
erated via products of the n choose 2 unique constituent for both gas- [24] and condensed phase systems [39–41].
pairwise polynomials of the higher-order terms. For exam- The computational efficiency of DFTB (i. e., 102–103 improve-
ple, a three-body term has 3 choose 2 ¼ 3 pairs, which ment over DFT) arises from its approximate Hamiltonian
yields the following expression for the ChIMES three-body and use of (generally) short-ranged empirical functions,
energy: though it still requires solving for electronic eigenstates,
� � � which generally scales as OðN3 Þ, similar to DFT calculations.
3
Ei1 i2 i3 ¼ f eci ei ri1 i2 f eci ei ri1 i3 f eci ei ri2 i3 �
1 2 1 3 2 3
Nevertheless, the computational efficiency of DFTB has
0 a two-fold immediate advantage: (1) simulations can be run
X
O3B X
O3B X
O3B
� � � (3)
C ei1 ei2 ei3
Tm s
ei1 ei2
Tp s
ei1 ei3
Tq s
ei2 ei3 for several orders of magnitude longer than what is feasible
mpq i 1 i2 i1 i3 i2 i 3
m¼0 p¼0 q¼0 with standard DFT-MD simulations (e. g., nanoseconds vs.
tens of picoseconds, e. g., relevant to high-pressure carbon
The above expression contains a triple sum for the condensation [42–44]), and (2) large numbers of statistically
i1 i2 ; i1 i3 ; i2 i3 polynomials over the hypercube up to order independent reactive MD simulations can be run simulta-
O3B , and include a single permutationally invariant co- neously, allowing for the generation of ensemble statistics.
efficient for each set of powers and atom types, Cempq i ei ei
. The 1 2 3
DFTB begins by assuming spherically symmetric charge
primed sum denotes that only terms for which two or more densities centered on atoms and then expanding the Kohn-
of the m; p; q polynomial powers are greater than zero are Sham DFT energy expression to second order in charge
included in order to guarantee that three distinct atom-cen- fluctuations [45]. Hamiltonian and overlap matrix elements
ters are evaluated. The expression for 3 Ei1 i2 i3 also contains are pre-tabulated from a two-center interaction potential

Propellants Explos. Pyrotech. 2022, 47, e202200001 (3 of 11) © 2022 Wiley-VCH GmbH
Machine-Learning a Solution for Reactive Atomistic Simulations of Energetic Materials

with use of a minimal local atomic basis set. The remaining treme conditions, where sudden pressurization and heating
double counting and ion-ion repulsion terms are then can cause transient metallization and different bonding ar-
grouped into a “repulsive energy”, which is generally fit to rangements can lead to significant changes in electronic
short-range empirical functional forms (usually spanning states and the resulting chemical reactivity [50].
bonded distances or those of the first solvation shell, only). In this respect, ChIMES presents a convenient tool for in-
The resulting energy expression can be written as follows: cluding additional many-body effects within ERep . This can
allow models to overcome deficiencies due to the reduced
EDFTB ¼ EBS þ ECoul þ ERep þ EDisp : (4) dimensionality of the DFTB approximate Hamiltonian and
minimal basis set by matching features such as atomic
Here, EBS is the band structure energy, determined from forces, system energies, and stress tensor terms. Briefly,
the occupied electronic eigenstates, ECoul is the charge ChIMES can be used for ERep determination by first creating
transfer term, and ERep is the previously described repulsive a DFTB repulsive energy-specific training set, where DFTB
energy. EDisp is an externally added dispersion energy that forces and stress tensor components with the repulsive en-
can be used to improve accuracy for weakly interacting ma- ergy set to zero are subtracted from their respective DFT
terials [46]. The empirical nature of ERep allows DFTB to be values, e. g.:
“tuned” to specific physical and/or chemical properties such
* ¼ Ft
FtRep FtQM;DFTBa
as isothermal compression curves or small cluster data, fre- a DFTa
i i i
(5)
quently computed from DFT or even higher-level quantum * ¼ st
chemical data. stRep aa DFTaa stQM;DFTBaa : (6)
DFTB models generally exist within a “sweet-spot” be- * ¼ Et
EtRep EtQM;DFTBaa : (7)
tween DFT and completely empirical MD models in terms aa DFTaa

of accuracy and ease of development. The quantum me-


chanical part of the calculation tends to be relatively accu- Here, t corresponds to a specific MD configuration, a to
rate for many conditions and materials, allowing for a rea- the cartesian direction, and i is the atomic index. The ‘*’ is
sonable representation of the system’s electronic ground used to denote that the quantities being computed are part
state [28]. This is especially true under extreme conditions, of the training set, and ‘QM,DFTB’ refers to the quantum
where electron repulsions tend to dominate interactions components of the DFTB calculation, i. e., only forces, en-
and exchange-correlation interactions do not need to be ergies, and stresses from EBS and ECoul :
determined as precisely [47]. In addition, the empirical ERep ChIMES parameters can then be determined from mini-
a function is only fit to very short-ranged interactions, mizing the following objective function:
which yields several advantages. Due to the underlying
vffiffiffiffiffiffi0
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffihffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiiffiffiffiffi1
ffiffiffiffi
quantum mechanics, DFTB models generally require far less u PM PN P3 2
u t t*
training data than full ML–IAPs, and generally exhibit a high u B t¼1 i¼1 a¼1 FChIMESa FRepa C
u B i i

C
degree of transferability across materials and thermody- u B h i2 C
u1 B P C
namic conditions [48]. For example, a recent model for bulk Fobj ¼u þ Mt¼1 EtChIMES EtRep * C; (8)
uN B B C
α-Ti as well as bulk TiH2 and its surface chemistry was de- u B d
C
u @ h i A
termined from several dozen MD snapshots performed on t PM P3 2
t t*
þ s ChIMESaa s Repaa
the TiH2 12 atom unit cell over a range of pressures [28]. In t¼1 a¼1

this regard, DFTB has potential as an intermediate model-


ing approach, where some initial investment in determining where M is the total number of configurations in the train-
an accurate ERep can result in a highly efficient model that ing set, and Nd is the total number of data entries (3MN
retains most of the accuracy of higher levels of theory. force components plus 3M stress tensor components).
These models can be used as a proxy for full DFT calcu- ChIMES/DFTB models have been created for several sys-
lations for both determinations of structural and chemical tems, including high-level quantum chemical organic clus-
information, as well as to generate significantly larger quan- ters [27], detonating energetic materials [26], and metallic
tities of training data to ultimately be used for the determi- systems and their surfaces [25, 28].
nation of inter-atomic potentials [26].
Consequently, the most significant difficulty with creat-
ing DFTB models frequently resides with a determination of 2.3 ChIMES Overview: Special Considerations
ERep . In the past, ERep has been determined based on a clus-
ter or high-symmetry crystalline data (e. g., cubic lattice Several practical considerations factor into the develop-
compression curves) to pair-wise functional forms, such as a ment of ML–IAPs. Perhaps most importantly, one must con-
simple polynomial power series or cubic spline [39]. How- sider whether the provided basis and training data are
ever, these approaches are unable to correct for the greater compatible with key governing physics in the target sys-
than two-center terms that can be significant for some sys- tem. For example, the standard ChIMES form (e. g., Eq’s 1–3
tems [49]. This is particularly true for materials under ex- above) combined with Newtonian molecular dynamics do

Propellants Explos. Pyrotech. 2022, 47, e202200001 (4 of 11) © 2022 Wiley-VCH GmbH
Research Article R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried

not explicitly account for spin, long-range charge inter- for agile training data generation that have been in-
actions, quantum nuclear vibrational effects, or variable corporated into the ChIMES framework are described.
electronic temperature. In many cases, these assumptions ChIMES training data is typically generated via DFT-MD.
are reasonable (e. g., in the cases of tightly coupled spin Models trained to these data will excel at describing favor-
state and ion configuration, and charge screening effects); able regions of physicochemical space but can struggle to
however, in cases where they are not, supplemental terms describe relatively unfavorable regions (i. e., corresponding
should be added [22, 51, 52]. to relatively rare events), which are likely to be under-repre-
Other considerations include the need to strike a bal- sented in the training set. For some materials and applica-
ance between accuracy and efficiency. In principle, increas- tions (e. g., non-reactive systems, solid-state materials, and
ing model complexity enables improved recovery of train- gas-phase chemistry), supplementary training config-
ing data (overfitting notwithstanding); however, doing so urations can be generated via straightforward means (e. g.,
concurrently increases computational expense. Identifying well-established enhanced sampling methods, random dis-
that balance is a non-trivial endeavor frequently performed placements, and manual perturbations to system lattice pa-
using methods like holdout cross-validation, but this can be rameters). However, generating suitable configurations for
difficult when trying to select from 103–105 parameters. To energetic materials is substantially more challenging since
overcome this challenge, advanced linear solvers are em- models must describe chemically reactive condensed phase
ployed, (e. g., LARS and LASSO) [53–55] which use absolute molecular systems characterized by disparate energy scales
value-based regularization to automatically set parameters (i. e., governing bonded, non-bonded, and reactive inter-
deemed inconsequential to zero; this has the effect of de- actions) and an expansive configurational space for the
creasing model size while increasing both stability and many molecular species formed.
computational efficiency. Note that these algorithms can be For relatively simple materials, iterative refinement can
distributed for application to large fitting problems (e. g., be used [22, 56] to generate the target configurations. In
104–105 parameters) [23]. this approach, an initial model is fit to an initial training set
(Figure 1, parts 1 and 2), then used to launch simulations
(Figure 1, part 3). If the resulting physical predictions are in-
2.4 ChIMES Overview: On Accuracy, Efficiency, sufficiently accurate, a set of configurations from the result-
Robustness, and Agility ing trajectories can then be combined with the initial train-
ing set, once DFT forces, energies, and stresses are assigned
As with any ML–IAP framework, ChIMES model develop- (Figure 1, parts 4 and 6). The updated training set can then
ment accuracy, efficiency, robustness, and agility is largely be used to train a new model (Figure 1 parts 1 and 2), and
governed by three interrelated factors: the cycle is repeated until the desired model accuracy is
* Model complexity achieved. Note that ChIMES is particularly well suited to this
* Model optimization approach since models can be generated rapidly. For more
* Training data generation efficiency complex fitting problems (e. g., higher chemical complex-
Broadly speaking, increasing model complexity (i. e., ity), more advanced approaches are needed. In these cases,
number of fitting parameters) enables improved model ac- active learning can greatly accelerate model generation, by
curacy; however, large models are computationally in- selecting specific configurations and/or molecular struc-
efficient and generally challenging to optimize due to the tures most likely to be underrepresented in the training set
expansive parameter space (this is especially true for non- (Figure 1. part 5) [57].
linear models). This is further complicated by the fact that
ML models “learn” these parameters solely from the avail-
able training data. Irrespective of model complexity, an in- 3 Results and Discussion
adequately large and/or informative training set will pre-
clude developing an accurate (and possibly, physically A central goal of ChIMES is enabling an agile and robust
reasonable) model. Ultimately, training data plays a critical framework for developing quantum-accurate models for EM
role in determining ML–IAP accuracy and robustness; how- capable of predicting material thermochemistry, the shock
ever, its generation remains a grand challenge, raising Hugoniot equation of state, and shockwave phenomena in-
questions including: cluding carbon coagulation and hot spot initiation. The ap-
* How can training data be generated in an efficient man- plication of ChIMES to systematically more complex materi-
ner? als and problems has simultaneously enabled evaluation of
* How can maximally informative training data be identi- the underlying framework efficacy and an improved under-
fied? standing of the chemistry driving detonation-induced ma-
In section 2.2 above, use of ChIMES-corrected DFTB as terial response.
an efficient proxy for DFT during training data generation In this section, an overview of some of these ChIMES
and/or labeling was discussed. Below, additional solutions modeling efforts is provided, including:
* Understanding shockwave-driven carbon condensation

Propellants Explos. Pyrotech. 2022, 47, e202200001 (5 of 11) © 2022 Wiley-VCH GmbH
Machine-Learning a Solution for Reactive Atomistic Simulations of Energetic Materials

Figure 1. The ChIMES active learning framework.

* Sensitization in liquid explosives oxygen purification kinetics) [33, 34]. Moreover, these simu-
* QM-accurate gas-phase chemistry calculations lations were validated through UF experiments, which veri-
fied the predicted cluster sizes, shape, chemical composi-
tion, and formation time scales (see Figure 2) [33]. This
3.1 Understanding Shockwave-Driven Carbon effort has provided a powerful validation of ChIMES and ap-
Condensation plication to chemically complicated problem spaces. This
effort has also served as a starting point to systematically
3.1.1 Shocked Carbon Monoxide investigate more complex problems, including carbon con-
densation in the hydrogen-free EM 3,4-bis(3-nitrofurazan-4-
Non-ideal EM can release up to 10 % of their detonation en- yl)furoxan (DNTF), discussed in greater detail below.
ergy through carbon condensation, which can have sig-
nificant implications for EM performance and sensitivity.
However, this complicated condensed-phase reaction-driv- 3.1.2 Shocked DNTF
en process remains poorly understood due to the asso-
ciated spatiotemporal scales (i. e., spanning 0.1 to 10s of ChIMES is also enabling improved understanding of EM for
nanometers and up to 100 s of nanoseconds), which have which experimental data are limited. For example, a
so far, precluded direct experimental observation. Though
considered rapid by experimental standards, these scales
are coincidentally at the upper limit of accessibility to atom-
istic simulation methods, requiring quantum accurate (de-
fined here as having an accuracy similar to Kohn-Sham DFT)
models capable of simulation on large scale. This process is
further complicated by high chemical complexity (i. e., the
manifold of chemistry that unravels as C/H/O/N-containing
materials begin to react under high T and P). To better un-
derstand this process, a reductionist investigation of carbon
condensation in shock compressed cryogenic liquid carbon
monoxide was conducted, through a hybrid experimental
and simulation study. In particular, ChIMES models [56] ca-
pable of simulation on the same time scales and over-
lapping length scales as those achievable through ultra-
fast-laser-driven shock (“UF”) experiments were developed.
These models were used to predict shock-induced Figure 2. Left: ChIMES simulations predicting nanocarbon for-
mation on experimentally relevant scales. Black and red beads are
nano-carbon formation over 10s of ps timescales, and pro-
carbon and oxygen atoms, respectively; atom connections are
vided an in-depth description of the overall system evolu- drawn as a visual aid and do not necessarily correspond to chem-
tion (e. g., cluster size distribution, composition, and classi- ical bonds. Right: TEM image from UF experiments. Red semicircles
cal governing growth kinetics) associated chemical kinetics indicate individual nanocarbon particles, for which a simulation-
(e. g., reactive transport of carbon into and between clus- predicted image is also shown in the zoomed-in dashed yellow out-
ters, the evolution of the fluid surrounding clusters, and line.

Propellants Explos. Pyrotech. 2022, 47, e202200001 (6 of 11) © 2022 Wiley-VCH GmbH
Research Article R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried

ChIMES-corrected DFTB model was developed for DNTF 6.6 km/s, i. e., closer to the experimentally reported range of
that was transferable to T/p states relevant to detonation, 7.1–7.6 km/s than the value predicted using standard DFTB
with the aim of shedding light on questions including: (1) models (6.0 km/s) [50], demonstrating that ChIMES can out-
what are the major chemical kinetic steps associated with perform the accuracy of semi-empirical quantum ap-
its shock decomposition? (2) what is its Hugoniot equation proaches. Moreover, the model is approximately 5 orders of
of state? (3) how do these properties compare with those magnitude more computationally efficient than DFT and
for other carbon-rich EM? and (4) what are the implications scales linearly with system size, enabling simulations for
for nanocarbon formation mechanisms in DNTF and other larger spatiotemporal scales, as was done in the presently
EM? For this study, a ChIMES-corrected DFTB model (i. e., described 1 ns MSST simulations of a 2048 atom system.
rather than a full-ChIMES model) was developed for en- Here, preliminary results for direct-shock simulations of
hanced agility, since the underlying semi-empirical quan- porous liquid HN3 using ChIMES are presented. Though
tum method significantly reduces the ChIMES model com- more computationally demanding than the previous MSST
plexity (and correspondingly, training set size) requirements simulations due to use of over 50-fold more atoms, it is a
to generate a high accuracy model. more powerful method for direct observation of phenom-
To address these questions, the ChIMES-corrected DFTB ena like hotspot formation. LAMMPS [62] with the ChIMES
model was leveraged in extended Lagrangian Multiscale calculator library [63] was used to simulate an orthogonal
Shock Technique (MSST) simulations [58–60], which con- computational cell with initial parameters a = 16.52 Å, b =
strain the sampled thermodynamic states to those for a 155.18 Å, and c = 622.77 Å and a total of 108,800 atoms
steady planar shockwave within continuum theory (i. e., (27,200 molecules). Periodic boundary conditions are ap-
simulating steady shock waves by constraining the stress plied in the x- and y-axis while in the z-axis where a shock-
and energy of an MD simulation to the Rayleigh line and wave is generated, a non-periodic boundary condition is
the Hugoniot energy relations). This simulation approach used. A fixed wall is placed at the left edge of the simu-
enabled estimation of the DFT Chapman-Jouguet (C J) det- lation box in the z-axis and a cylindrical pore with the radi-
onation state, Hugoniot curve, and shockwave-driven us r = 40 Å is created 200 Å from the fixed wall. An initial
chemistry [26]. On DFTB-accessible scales, DNTF chemical velocity -up is provided for all atoms in the NVE simulation.
evolution was found to be largely characterized by release The particles impact the wall and reflect, generating a
of CO, CO2, and N2, and in contrast to oxygen-rich and oxy- shockwave that propagates along the z-axis.
gen-balanced materials [42, 61], NO3 does not form in sig- Figure 3 shows the evolution of system density as a
nificant quantities nor appear to inhibit chemical evolution. shockwave passes through the material for the simulation
Formation and slow evolution of large CNO-containing spe- with up = 1.5 km/s. After about 7.5 ps, the pore has col-
cies (i. e., which are likely precursors to the experimentally lapsed, and hotspot formation begins. As can be seen in
observed carbon condensates) were also found to be a Figure 3, density at hot spot location exceeds that of both
characteristic feature of shock-driven DNTF chemistry. Ulti- the shocked and unshocked material. As the hot spot re-
mately, this combined ChIMES/DFTB/MSST simulation ap- laxes, rarefaction waves are observed, momentarily creating
proach provides a means of reaching far longer timescales a region near the fixed wall that has lower density than that
with “quantum accuracy.” Beyond accelerating first-princi- of the shocked regions. The Shock Trapping Internal Boun-
ples-based EM simulation studies, this approach can also daries (STIBs) method [6] was used to study the evolution of
significantly expedite development efforts for stand-alone the hot spot, in which small immobile material regions are
ChIMES ML–IAP models, by serving as an efficient, effective imposed between the shockwave and the hot spot. These
means of generating DFT-quality training and validation immobile STIBs regions have the widths of 20 Å, which is
data for materials where experimental data are scarce. wider than the cutoff distance for force interactions be-
tween particles (8 Å in this case); atoms outside of these re-
gions (i. e., the hot spot and its surrounding material is in
3.2 Sensitization in Liquid Explosives between the STIBs) are fully mobile. With these constraints,
atoms of interest interact with each other and with the im-
More recently, ChIMES models have been created to study mobile STIB atoms, allowing for direct study of the time
shocked liquid hydrazoic acid (HN3, i. e., the simplest co- evolution of the hot spot. Material outside two STIBs re-
valent azide-based EM) for which previous DFTB MSST sim- gions are neglected to reduce computational cost. Note
ulations suggested rapid detonation chemistry [50]. In a that the STIBs method must be applied when all shock-
previous study [23], a ChIMES model capable of accurately waves are sufficiently far from the area of interest, which in
reproducing DFT-predicted material structure, dynamical this case, it is at the simulation time of 10 ps.
properties, and chemistry for a wide range of unreactive The evolution of the system temperature in the STIBs
and decomposition conditions of liquid HN3 was developed. simulation for porous HN3 is shown in the right panel of
This model was then leveraged in MSST simulations to Figure 4. For comparison, a similar simulation for liquid HN3
study shockwave-driven detonation of liquid HN3. ChIMES was performed with no initial void, and the result is also
predicted a C J detonation velocity between 6.5 and displayed in Figure 4 (left panel). For porous HN3, the pore

Propellants Explos. Pyrotech. 2022, 47, e202200001 (7 of 11) © 2022 Wiley-VCH GmbH
Machine-Learning a Solution for Reactive Atomistic Simulations of Energetic Materials

simulation of the liquid HN3 (without pore collapse) where


the system temperature is nearly uniform over the whole
material. In both cases, the system temperature increases in
time as chemical reactions occur.
These preliminary results indicate that the presence of a
pore significantly affects the evolution of the shocked sys-
tem. Overall, this study shows the ability of ChIMES simu-
lations to provide quantum accurate results for shocked
materials that would not be possible using quantum me-
chanical calculation directly.

3.3 Quantum-Accurate Gas-Phase Chemistry Calculations

Recently, a DFTB/ChIMES model was developed for organic


materials, by training to a small subset (i. e., only ~ 0.3 %) of
the ANI-1x dataset, which contains a diverse set of molec-
ular conformation at equilibrium and non-equilibrium geo-
metries, with forces and energies assigned via hybrid DFT
[35]. As described in section 2.2, the benefit of using
Figure 3. Time evolution (i. e., from top to bottom in 2.5 ps incre- ChIMES in this manner is that many-body interactions can
ments) of a 108,800 atom HN3 system containing a cylindrical void be corrected for in a systematic and rapidly tunable process
shocked via impact with a piston to the left of the system. System by varying the ChIMES polynomial order parameters. Addi-
density is indicated in g cm 3 by colors in the color bar. At 10 ps, tionally, the underlying semi-empirical framework enables
the STIBs method was applied to the area of interest (i. e., contain- the determination of the electronic states of a system, al-
ing the hot spot), between the two vertical red lines shown in the
lowing for possible thermal electronic excitations, which
bottom panel.
can be significant for many energetic materials under shock
compression and can lower activation energies for re-
activity.
The performance of the DFTB/ChIMES model in predict-
ing energies, forces and vibrational frequencies has also
been tested on many additional quantum datasets. Ulti-
mately, this model yielded strong agreement with reference
data from either hybrid DFT, coupled-cluster calculations or
experiments. While all training data for developing DFTB/
ChIMES were taken from gas-phase clusters only, this model
has also been shown to reproduce the experimental struc-
ture of graphite (a well-known challenge for DFT).
Here, new results for the enthalpy of formation (ΔHf) of
EM (Figure 5) predicted by standard DFTB (with the 3ob pa-
rameter set) and a DFTB/ChIMES model developed in [64],
using protocols described in [27, 64] are presented. As
shown in [64], the highly accurate ccCA-PS3 method yields
excellent agreement with available experimental data, with
a mean absolute error (MAE) of ~ 12,500 J/mol. DFTB pre-
Figure 4. Time evolution (i. e., from top to bottom in 25 ps incre- dicts ΔHf in good agreement with experimental data for
ments) of the temperature in the STIBs simulations for pure liquid some cases (for example RDX and HMX). DFTB/ChIMES
HN3 (left panel) and porous HN3 (right panel). System temperature shows an improvement over DFTB in all cases considered
is indicated in K by colors in the color bar. here. The MAE values for DFTB and DFTB/ChIMES are
~ 42,000 and 25,000 J/mol, respectively. For the molecules
where experimental data is not available (NTO, DADE and
collapse creates a hot spot with the temperature in some LLM-105), DFTB/ChIMES predicts the values closer to that of
regions > 2000 K, while the temperature of the shocked ccCA-PS3 than DFTB. It is worth mentioning that the accu-
material is only ~ 750 K. At later times, the hot spot diffuses racy of the DFTB/ChIMES model is very similar to G3 and G4
its heat to surrounding atoms and the system temperature quantum composite methods used in [64], although it uses
becomes uniform after 50 ps. This is in contrast with the only a fraction of computational resources required by G3

Propellants Explos. Pyrotech. 2022, 47, e202200001 (8 of 11) © 2022 Wiley-VCH GmbH
Research Article R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried

work is helpful in this regard, but additional development is


needed for full generalization to arbitrary materials.
A second challenge is that achieving quantum-level ac-
curacy for highly complex chemistry may require the addi-
tion of explicit � 5-body interactions in the potential energy
function, whereas ChIMES applications to date have in-
cluded a maximum of 4-body terms. This is a straightfor-
ward generalization of the models presented here but
would represent a significant increase in the number of pa-
rameters. Handling a very large number of possible poten-
tial parameters will require very scalable linear regularized
optimization and parameter selection implementations
(such as LASSO), and highly effective methods to generate
training configurations.
Finally, the ChIMES approach does not currently treat
some quantum aspects of material behavior such as elec-
tron spin and electron temperature. Nuclear quantum vibra-
Figure 5. Comparison of predicted enthalpies of formation of some tional effects could be treated through path-integral molec-
energetic molecules for DFTB, DFTB/ChIMES, ccCA-PS3 (ref. 64) with ular dynamics [65] or quantum thermostats [66], but have
experimental data (ref. 67) and predictions from an atom-equiv- not yet been investigated with the ChIMES framework.
alent method based on quantum chemical calculations at the
B3LYP/6-31G* level (Rice et al., ref. 68).
Acknowledgements

or G4. For example, PETN quantum composite calculations, This work was performed under the auspices of the U.S. Depart-
which involve geometry optimization and vibrational fre- ment of Energy by Lawrence Livermore National Laboratory under
Contract No. DE-AC52-07NA27344. LLNL-JRNL-830464.
quency generation for 100s of trial configurations, G3 and
G4 requires approximately 500 GB of physical memory dis-
tributed among 32 CPUs, for 21 days. In contrast, DFTB/
ChIMES calculations including geometry optimization and Data Availability Statement
frequency determination use a single node and take about
30 minutes. The data that support reported findings are available within
all cited references and/or the manuscript body.

4 Conclusion
References
There is clearly a need for highly accurate and scalable
atomistic simulations of EM spanning temperatures from [1] M. R. Manaa, L. E. Fried, The reactivity of energetic materials
~ 300 to ~ 5000 K and pressures from 0 to ~ 50,000 MPa. under high pressure and temperature, in Advances in Quantum
Across this range of conditions, there are tremendous Chemistry (Ed. J. Sabin), Academic Press, London 2014, p. 221.
[2] A. C. Landerville, M. W. Conroy, M. M. Budzevich, Y. Lin, C. T.
changes in speciation and chemical reactivity. The ChIMES
White, I. I. Oleynik, Equations of state for energetic materials
ML–IAP provides a promising path forward and has been from density functional theory with van der Waals, thermal,
successfully applied to a number of systems including mod- and zero-point energy corrections. Appl. Phys. Lett. 2010, 97,
el energetic materials. Within the ChIMES family, DFTB/ 251908.
ChIMES ML–IAPs can be generated most rapidly and require [3] R. Menikoff, T. D. Sewell, Constituent properties of HMX need-
minimal training sets, while the stand-alone ChIMES ML– ed for mesoscale simulations. Combust. Theory Model. 2002, 6,
IAPs are scalable to millions of atoms. Further development 103.
of ChIMES is required to access the full range of species and [4] Q. An, S. V. Zybin, W. A. Goddard, A. Jaramillo-Botero, M. Blan-
states typical of reacting CHNO energetic materials, such as co, S. N. Luo, Elucidation of the dynamics for hot-spot ini-
tiation at nonuniform interfaces of highly shocked materials.
RDX or TATB, which is the subject of ongoing work.
Phys. Rev. B – Condens. Matter Mater. Phys. 2011, 84, 220101.
A key challenge common to all ML–IAP development ef- [5] T. R. Shan, R. R. Wixom, A. P. Thompson, Extended asymmetric
forts is that the DFT-MD simulations typically used to gen- hot region formation due to shockwave interactions following
erate training data over-sample statistically favorable con- void collapse in shocked high explosive. Phys. Rev. B 2016, 94,
figurations while unfavorable configurations important for 054308.
chemistry, such as transition states, are under-sampled.
Data selection through the ChIMES active learning frame-

Propellants Explos. Pyrotech. 2022, 47, e202200001 (9 of 11) © 2022 Wiley-VCH GmbH
Machine-Learning a Solution for Reactive Atomistic Simulations of Energetic Materials

[6] B. W. Hamilton, M. P. Kroonblawd, C. Li, A. Strachan, A Hot- [24] M. Stöhr, L. Medrano Sandonas, A. Tkatchenko, Accurate Many-
spot’s Better Half: Non-Equilibrium Intra-Molecular Strain in Body Repulsive Potentials for Density-Functional Tight Binding
Shock Physics. J. Phys. Chem. Lett. 2021, 12, 2756. from Deep Tensor Neural Networks. J. Phys. Chem. Lett. 2020,
[7] J. M. Winey, Y. Toyoda, Y. M. Gupta, Near-optimal combination 11, 6835.
of high performance and insensitivity in a shock compressed [25] N. Goldman, B. Aradi, R. K. Lindsey, L. E. Fried, Development of
high explosive single crystal. J. Appl. Phys. 2021, 130, 015902. a Multicenter Density Functional Tight Binding Model for Plu-
[8] B. C. Barnes, J. K. Brennan, E. F. C. Byrd, S. Izvekov, J. P. Lar- tonium Surface Hydriding. J. Chem. Theory Comput. 2018, 14,
entzos, B. M. Rice, Toward a Predictive Hierarchical Multiscale 2652.
Modeling Approach for Energetic Materials, in Challenges and [26] R. K. Lindsey, S. Bastea, N. Goldman, L. E. Fried, Investigating
Advances in Computational Chemistry and Physics (Ed. J. Leszc- 3,4-bis(3-nitrofurazan-4-yl)furoxan detonation with a rapidly
zynski), Springer, Cham, 2019 p. 229. tuned density functional tight binding model. J. Chem. Phys.
[9] W. Kohn, L. J. Sham, Self-consistent equations including ex- 2021, 154, 164115.
change and correlation effects. Phys. Rev. 1965 140, A1133. [27] C. Huy Pham, R. K. Lindsey, L. E. Fried, N. Goldman, High Accu-
[10] A. Strachan, E. M. Kober, A. C. T. Van Duin, J. Oxgaard, W. A. racy Semi-Empirical Quantum Models Based on a Reduced
Goddard, Thermal decomposition of RDX from reactive molec- Training Set. ChemRXIV, 2021, doi: 10.26434/chemrxiv-2021-
ular dynamics. J. Chem. Phys. 2005, 122, 054502. 2clcb.
[11] D. Guo, S. V. Zybin, A. An, W. A. Goddard, F. Huang, Prediction [28] N. Goldman, K. E. Kweon, B. Sadigh, T. W. Heo, R. K. Lindsey,
of the Chapman-Jouguet chemical equilibrium state in a deto-
C. H. Pham, L. E. Fried, B. Aradi, K. Holliday, J. R. Jeffries, B. C.
nation wave from first principles based reactive molecular dy-
Wood, Semi-Automated Creation of Density Functional Tight
namics. Phys. Chem. Chem. Phys. 2016, 18, 2015.
Binding Models through Leveraging Chebyshev Polynomial-
[12] D. Bedrov, G. D. Smith, T. D. Sewell, Temperature-dependent
Based Force Fields. J. Chem. Theory Comput. 2021, 17, 4435.
shear viscosity coefficient of octahydro-1,3,5,7-tetranitro-
[29] P. Yoo, M. Sakano, S. Desai, M. M. Islam, P. Liao, A. Strachan, A
1,3,5,7-tetrazocine (HMX): A molecular dynamics simulation
Neural network reactive force field for C, H, N, and O systems.
study. J. Chem. Phys. 2000, 112, 7203.
npj Comput. Mater. 2021, 7, 1.
[13] M. P. Kroonblawd, and T. D. Sewell, Theoretical determination
of anisotropic thermal conductivity for crystalline 1,3,5-tri- [30] M. A. Cusentino, M. A. Wood, A. P. Thompson, Explicit Multiele-
amino-2,4,6-trinitrobenzene (TATB). J. Chem. Phys. 2013 139, ment Extension of the Spectral Neighbor Analysis Potential for
074503. Chemically Complex Systems. J. Phys. Chem. A 2020, 124, 5456.
[14] A. C. T. Van Duin, S. Dasgupta, F. Lorant, W. A. Goddard, Re- [31] L. Zhang, J. Han, H. Wang, R. Car, E. Weinan, Deep Potential
axFF: A Reactive Force Field for Hydrocarbons. J. Phys. Chem. A Molecular Dynamics: A Scalable Model with the Accuracy of
2001, 105, 9396. Quantum Mechanics. Phys. Rev. Lett. 2018, 120, 143001.
[15] B. A. Steele, S. M. Clarke, M. P. Kroonblawd, M. P., I. F. W. Kuo, [32] E. J. Reed, M. R. Manaa, L. E. Fried, K. R. Glaesemann, J. D. Joan-
P. F. Pagoria, S. N. Tkachev, J. S. Smith, S. Bastea, L. E. Fried, nopoulos, A transient semimetallic layer in detonating nitro-
J. M. Zaug, E. Stavrou, O. Tschauner, Pressure-induced phase methane. Nat. Phys. 2008, 4, 72.
transition in 1,3,5-triamino-2,4,6-trinitrobenzene (TATB). Appl. [33] M. R. Armstrong, R. K. Lindsey, N. Goldman, M. H. Nielsen, E.
Phys. Lett. 2019, 114, 191901. Stavrou, L. E. Fried, J. M. Zaug, S. Bastea, Ultrafast shock syn-
[16] W. P. Bassett, B. P. Johnson, N. K. Neelakantan, K. S. Suslick, thesis of nanocarbon from a liquid precursor. Nat. Commun.
D. D. Dlott, Shock initiation of explosives: High temperature 2020 111, 1.
hot spots explained. Appl. Phys. Lett. 2017, 111, 061902. [34] R. K. Lindsey, N. Goldman, L. E. Fried, S. Bastea, Chemistry-
[17] E. B. Watkins, K. A. Velizhanin, D. M. Dattelbaum, R. L. Gus- Mediated Ostwald Ripening in Carbon-Rich C/O Systems at Ex-
tavsen, T. D. Aslam, D. W. Podlesak, R. C. Huber, M. A. Firestone, treme Conditions, Nat. Commun. 2022, 13, 1424.
B. S. Ringstrand, T. M. Willey, M. Bagge-Hansen, R. Hodgin, L. [35] J. S. Smith, O. Isayev, A. E. Roitberg, ANI-1: an extensible neural
Lauderbach, T. Van Buuren, N. Sinclair, P. A. Rigg, S. Seifert, T. network potential with DFT accuracy at force field computa-
Gog, Evolution of Carbon Clusters in the Detonation Products tional cost. Chem. Sci. 2017, 8, 3192.
of the Triaminotrinitrobenzene (TATB)-Based Explosive PBX [36] W. Wang, B. C. Shepler, B. J. Braams, J. M. Bowman, Full-dimen-
9502. J. Phys. Chem. C 2017, 121, 23129. sional, ab initio potential energy and dipole moment surfaces
[18] F. Emmert-Streib, Z. Yang, H. Feng, S. Tripathi, M. Dehmer, An for water. J. Chem. Phys. 2009, 131, 054511.
Introductory Review of Deep Learning for Prediction Models [37] J. Tersoff, Empirical interatomic potential for carbon, with ap-
With Big Data. Front. Artif. Intell. 2020, 3, p. 4. plications to amorphous carbon. Phys. Rev. Lett. 1988, 61, 2879.
[19] F. Ercolesi, J. B. Adams, Interatomic potentials from first-princi- [38] B. Hourahine, B. Aradi, V. Blum, F. Bonafé, A. Buccheri, C. Ca-
ples calculations: The force-matching method. EPL 1994, 26, macho, C. Cevallos, M. Y. Deshaye, T. Dumitric, A. Dominguez,
583. S. Ehlert, M. Elstner, T. Van Der Heide, J. Hermann, S. Irle, J. J.
[20] V. L. Deringer, A. P. Bartók, N. Bernstein, D. M. Wilkins, M. Cer- Kranz, C. Köhler, Y. Kowalczyk, T. Kubař, I. S. Lee, V. Lutsker, R. J.
iotti, G. Csányi, Gaussian Process Regression for Materials and Maurer, S. K. Min, I. Mitchell, C. Negre, T. A. Niehaus, A. M. N.
Molecules. Chem. Rev. 2021, 121, 10073. Niklasson, A. J. Page, A. Pecchia, G. Penazzi, M. P. Persson, J.
[21] R. K. Lindsey, L. E. Fried, N. Goldman, ChIMES: A Force Matched Řezáč, C. G. Sánchez, M. Sternberg, M. Stöhr, F. Stuckenberg, A.
Potential with Explicit Three-Body Interactions for Molten Car- Tkatchenko, V. W. Z. Yu, T. Frauenheim, DFTB + , a software
bon. J. Chem. Theory Comput. 2017, 13, 6222. package for efficient approximate density functional theory
[22] R. K. Lindsey, L. E. Fried, N. Goldman, Application of the
based atomistic simulations. J. Chem. Phys. 2020, 152, 124101.
ChIMES Force Field to Nonreactive Molecular Systems: Water
[39] A. K. Akshay, E. Wadbro, C. Köhler, P. Mitev, P. Broqvist, J.
at Ambient Conditions. J. Chem. Theory Comput. 2019, 15, 436.
Kullgren, CCS: A software framework to generate two-body
[23] C. H. Pham, R. K. Lindsey, L. E. Fried, N. Goldman, Calculation of
potentials using Curvature Constrained Splines. Comput. Phys.
the detonation state of HN3 with quantum accuracy. J. Chem.
Commun. 2021, 258, 107602.
Phys. 2020, 153, 224102.

Propellants Explos. Pyrotech. 2022, 47, e202200001 (10 of 11) © 2022 Wiley-VCH GmbH
Research Article R. K. Lindsey, C. Huy Pham, N. Goldman, S. Bastea, L. E. Fried

[40] M. P. Lourenço, E. C. Dos Santos, L. G. M. Pettersson, H. A. [54] B. Efron, T. Hastie, I. Johnstone, R. Tibshirani, H. Ishwaran, K.
Duarte, Accurate SCC-DFTB Parametrization for Bulk Water. J. Knight, J. M. Loubes, P. Massart, D. Madigan, G. Ridgeway, S.
Chem. Theory Comput. 2020, 16, 1768. Rosset, J. I. Zhu, R. A. Stine, B. A. Turlach, S. Weisberg, I. John-
[41] G. Dolgonos, B. Aradi, N. H. Moreira, T. Frauenheim, An im- stone, R. Tibshirani, Least angle regression. Ann. Stat. 2004, 32,
proved self-consistent-charge density-functional tight-binding 407.
(SCC-DFTB) set of parameters for simulation of bulk and mo- [55] J. Friedman, T. Hastie, R. Tibshirani, Regularization paths for
lecular systems involving titanium. J. Chem. Theory Comput. generalized linear models via coordinate descent. J. Stat.
2010, 6, 266. Softw. 2010, 33, 1.
[42] M. R. Manaa, E. J. Reed, L. E. Fried, N. Goldman, Nitrogen-Rich [56] R. K. Lindsey, N. Goldman, L. E. Fried, S. Bastea, Many-body re-
Heterocycles as Reactivity Retardants in Shocked Insensitive active force field development for carbon condensation in C/O
Explosives. J. Am. Chem. Soc. 2009, 131, 5483. systems under extreme conditions. J. Chem. Phys. 2020, 153,
[43] M. P. Kroonblawd, R. K. Lindsey, N. Goldman, Synthesis of func- 054103.
tionalized nitrogen-containing polycyclic aromatic hydro- [57] R. K. Lindsey, L. E. Fried, N. Goldman, S. Bastea, Active learning
carbons and other prebiotic compounds in impacting glycine for robust, high-complexity reactive atomistic simulations. J.
solutions. Chem. Sci. 2019, 10, 6091. Chem. Phys. 2020, 153, 134117.
[44] C. B. Cannella, N. Goldman, Carbyne Fiber Synthesis within [58] E. J. Reed, L. E. Fried, J. D. Joannopoulos, A Method for Tract-
Evaporating Metallic Liquid Carbon. J. Phys. Chem. C 2015, 119 able Dynamical Studies of Single and Double Shock Com-
(37), 21605. pression. Phys. Rev. Lett. 2003, 90, 4.
[45] M. Elstner, D. Porezag, G. Jungnickel, J. Elsner, M. Haugk, T. [59] E. J. Reed, L. E. Fried, M. R. Manaa, J. D. Joannopoulos, A Multi-
Frauenheim, Self-consistent-charge density-functional tight- Scale Approach to Molecular Dynamics Simulations of Shock
binding method for simulations of complex materials proper- Waves, in Chemistry at Extreme Conditions (ed. Riad Manaa, M.),
ties. Phys. Rev. B - Condens. Matter Mater. Phys. 1998, 58, 7260. Elsevier, 2005, p. 297.
[46] E. Caldeweyher, S. Ehlert, A. Hansen, H. Neugebauer, S. Spich- [60] E. J. Reed, L. E. Fried, W. D. Henshaw, C. M. Tarver, Analysis of
er, C. Bannwarth, S. Grimme, A generally applicable atomic- simulation technique for steady shock waves in materials with
charge dependent London dispersion correction. J. Chem. analytical equations of state. Phys. Rev. E – Stat. Nonlinear, Soft
Phys. 2019, 150, 154122. Matter Phys. 2006, 74, 056706.
[47] N. Goldman, E. J. Reed, I. F. W. Kuo, L. E. Fried, C. J. Mundy, A. [61] N. Goldman, S. Bastea, Nitrogen oxides as a chemistry trap in
Curioni, Ab initio simulation of the equation of state and ki- detonating oxygen-rich materials. J. Phys. Chem. A 2014, 118,
netics of shocked water. J. Chem. Phys. 2009, 130, 124517. 2897.
[48] N. Goldman, L. E. Fried, L. Koziol, Using Force-Matched Poten- [62] S. Plimpton (1995) Fast parallel algorithms for short-range mo-
tials To Improve the Accuracy of Density Functional Tight lecular dynamics. J. Comput. Phys. 117, 1, 1–19.
Binding for Reactive Conditions. J. Chem. Theory Comput. 2015, [63] rk-lindsey/chimes_calculator: Tools to interface ChIMES with
11, 4530. various external codes. https://github.com/rk-lindsey/chimes_
[49] N. Goldman, S. Goverapet Srinivasan, S. Hamel, L. E. Fried, M. calculator.
Gaus, M. Elstner, Determination of a density functional tight [64] M. R. Manaa, L. E. Fried, I. F. W. Kuo, Determination of en-
binding model with an extended basis set and three-body re- thalpies of formation of energetic molecules with composite
pulsion for carbon under extreme pressures and temperatures. quantum chemical methods. Chem. Phys. Lett. 2016, 648, 31.
J. Phys. Chem. C 2013, 117, 7885. [65] D. Marx, and M. Parrinello, Ab initio path integral molecular
[50] E. J. Reed, A. W. Rodriguez, M. R. Manaa, L. E. Fried, C. M. Tarv- dynamics: Basic ideas. J. Chem. Phys. 1996, 104, 4077.
er, Ultrafast detonation of hydrazoic acid (HN3). Phys. Rev. Lett. [66] T. E. Markland, M. Ceriotti, Nuclear quantum effects enter the
2012, 109, 038301. mainstream. Nat. Rev. Chem. 2018, 2, 1.
[51] L. Koziol, L. E. Fried, N. Goldman, Using force matching to de- [67] V. I. Pepkin, Y. N. Matyushin, Y. A. Lebdev, BACCA 1974, 23,
termine reactive force fields for water under extreme thermo- 1707.
dynamic conditions. J. Chem. Theory Comput. 2017, 13, 135. [68] B. M. Rice, S. V. Pai, J. Hare, Predicting heats of formation of
[52] O. T. Unke, S. Chmiela, M. Gastegger, K. T. Schütt, H. E. Sauce- energetic materials using quantum mechanical calculations.
da, K.-R. Müller, SpookyNet: Learning Force Fields with Elec- Combust. Flame 1999, 118, 445.
tronic Degrees of Freedom and Nonlocal Effects. Nat. Commun.
2021, 12, 7273. Manuscript received: January 1, 2022
[53] R. Tibshirani, Regression Shrinkage and Selection Via the Lasso. Revised manuscript received: February 16, 2022
J. R. Stat. Soc. Ser. B 1996, 58, 267. Version of record online: March 22, 2022

Propellants Explos. Pyrotech. 2022, 47, e202200001 (11 of 11) © 2022 Wiley-VCH GmbH

You might also like