Sunteți pe pagina 1din 61

Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article
Nanoindentation of the Triclinic Molecular Crystal 1,3,5-
Triamino-2,4,6-Trinitrobenzene: A Molecular Dynamics Study
Nithin Mathew, and Thomas D. Sewell
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b01103 • Publication Date (Web): 28 Mar 2016
Downloaded from http://pubs.acs.org on April 4, 2016

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society.


1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
Nanoindentation of the Triclinic Molecular Crystal
9
10
11
12 1,3,5-Triamino-2,4,6-Trinitrobenzene: A Molecular
13
14
15
16 Dynamics Study
17
18
19
20
21
22 Nithin Mathew and Thomas D. Sewell*
23
24
25
26
27
Department of Chemistry, University of Missouri-Columbia, Columbia, Missouri 65211-7600, USA
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
*Author to whom correspondence should be addressed. Electronic mail: sewellt@missouri.edu;
57 Telephone: +1 573-882-7725.
58
59
60 1
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 61

1
2
3 Abstract
4
5
6
Nanoindentation of the insensitive energetic molecular crystal 1,3,5-triamino-2,4,6-trinitrobenzene
7
8 (TATB) was studied using constant-temperature and constant-energy molecular dynamics simulations.
9
10 Displacement-controlled indentations at constant velocity were performed using a rigid, spherical indenter
11
12 on the three principal crystallographic planes—(100), (010), and (001). The force-displacement curve for
13
14 the (001) (basal) plane exhibits a distinct elastic region in agreement with the analytical solution for
15
16 indentation of an anisotropic half-space by a parabola of revolution. Stiffening precedes inelastic
17
18
19
deformation, and the elastic-inelastic transition occurs with kinking and delamination of the layered basal
20
21 planes and significant pile-up. The predicted nanoindentation hardness on the basal plane is 1.02±0.09
22
23 GPa. Nanoindentation on the (100) and (010) (non-basal) planes yields a non-Hertzian response; this is
24
25 attributed to an effective “softening” due to elastic bending of the molecular layers. Significant heating
26
27 occurs in the vicinity of the indenter during indentation with a higher temperature rise for the basal plane.
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 2
ACS Paragon Plus Environment
Page 3 of 61 The Journal of Physical Chemistry

1
2
3 1. Introduction
4
5
6
7
8 TATB (1,3,5-triamino-2,4,6-trinitrobenzene) is a highly insensitive energetic molecular crystal. The
9
10 molecules have the chemical formula C6 H 6 N 6O6 (see Figure 1a) and exhibit intramolecular hydrogen
11
12
bonding between O and H atoms on the adjacent nitro and amino groups. The only confirmed polymorph
13
14
15 exists in a triclinic crystal structure1 (space group P1 , see Figure 1b) although multiple computational2
16
17 and experimental3-5 studies have suggested the existence of other polymorphs. The crystal structure of
18
19 TATB exhibits a ‘graphitic-like’ packing motif, wherein nearly planar molecules form essentially two-
20
21
dimensional hydrogen-bonded sheets (layers) that stack in a skewed hexagonal-like arrangement with
22
23
24 only weak van der Waals bonding between the layers (see Figure 1c). This structural anisotropy results in
25
26 significant anisotropy in thermal,3-10 mechanical,11-17 and optical1 properties. Figure 1d shows the polar-
27
28 plot representation of the orientation-dependent Young’s modulus (see, e.g., ref. 18) of TATB calculated
29
30 using elastic coefficients at 0 K and 1 atm.19 The pronounced mechanical anisotropy of TATB is clearly
31
32 evident in Figure 1d, where the Young’s modulus values for the [100] and [010] directions (which
33
34
nominally correspond to the a and b lattice vectors, respectively) are much larger than that for the [001]
35
36
37 direction (which nominally corresponds to the c lattice vector).
38
39
40 An understanding of defects and deformation mechanisms in TATB is still at an early stage of
41
42 development. Experimental studies have reported high defect density in grown crystals of TATB20-22 and
43
44 ‘subtle structural phase transitions’5 under hydrostatic compression. Density functional theory (DFT)
45
46 electronic structure studies have predicted changes in hydrogen bonding13 and molecular rearrangement14
47
48 during hydrostatic compression. Generalized stacking-fault energy calculations, using all-atom
49
50
simulations with a classical force field, predicted very low unstable and stable stacking-fault energies on
51
52
53 the basal plane of TATB, suggesting easy dislocation glide.19
54
55
56
57
58
59
60 3
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 61

1
2
3 Material defects and inelastic/plastic deformation mechanisms are believed to influence energy
4
5
6
localization in solid explosives – often referred to as hot spots23-30 in the literature – and thus affect the
7
8 sensitivity of energetic molecular crystals. Specific mechanisms that are believed to mediate initiation of
9
10 detonation include (but are not limited to) void collapse,24 enhanced anharmonicity and multi-phonon up-
11
12 pumping31 (e.g., at defect sites), localized heating during plastic deformation,24,32-34 dislocation pile-
13
14 up,35,36 and bond-breaking due to hindered plasticity.37,38 Therefore, an understanding of defect structures
15
16 and deformation mechanisms can provide insight into molecular origins of energy localization
17
18
19 phenomena and anisotropic sensitivity in energetic molecular crystals such as TATB.
20
21
22 Nanoindentation has been used in multiple studies to systematically investigate deformation
23
24 mechanisms in molecular crystals.39-51 The low loads and small indentation depths in nanoindentation
25
26 facilitate the study of anisotropic mechanical response of molecular crystals, which are, in general,
27
28 elastically compliant and plastically hard. For brevity, the scope of the following discussion is limited to
29
30 nanoindentation in energetic molecular crystals. Ramos et al.42 studied the elastic-plastic response of
31
32 oriented single crystals of α-1,3,5-trinitroperhydro-1,3,5-triazine (α-RDX) using nanoindentation
33
34
35 experiments with a conical indenter. Despite the significant elastic anisotropy, the nanoindentation
36
37 hardness values of α-RDX, measured for different crystallographic planes, were found to be isotropic,
38
39 suggesting limited plasticity. These observations were supported by recent nanoindentation studies on
40
41 single crystals of α-RDX using a Berkovich indenter,50 although higher values were reported for the
42
43 reduced elastic modulus and hardness of the (210) face compared to the values reported in ref. 42. A
44
45
follow-up study by Ramos et al.44 showed that solution-washed (001) planes of α-RDX sustained the
46
47
48 highest values of the maximum shear stress, before the onset of plasticity during nanoindentation,
49
50 compared to thermally or mechanically cleaved planes. Hudson et al.52 performed nanoindentation
51
52 experiments with a Berkovich indenter on α-RDX crystals from different manufacturers and observed that
53
54 crystals with larger defect density have lower elastic moduli and higher shock sensitivity. Li et al.43
55
56 reported nanoindentation experiments on oriented single crystals of β-octahydro-1,3,5,7-tetranitro-
57
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 61 The Journal of Physical Chemistry

1
2
3 1,3,5,7-tetrazocine (β-HMX), using a Berkovich indenter. (The space group setting was not mentioned in
4
5
6
ref. 43, but the description of elastic moduli indicates that the P21/n setting was used.) They observed
7
8 ‘pop-in’ events during loading and nearly isotropic values of elastic modulus and nanoindentation
9
10 hardness for the two faces studied, namely (010) and another that was unspecified except as “the face
11
12 which is perpendicular to” the (010) face. Nanoindentation experiments by Kucheyev et al.51 on the (010)
13
14 face of β-HMX (P21/n space group setting) and LLM-105 (2,6-diamino-3,5-dinitropyrazine-1-oxide),
15
16 using both Berkovich and spherical indenters, showed fracture along the (011) cleavage plane in both
17
18
19
materials. The hardness value obtained by Kucheyev et al. for the (010) face of β-HMX was significantly
20
21 lower than the value reported in ref. 43.
22
23
24 Molecular dynamics (MD) simulations of nanoindentation, though performed at much higher
25
26 loading rates and smaller indentation depths compared to experimental studies, can provide useful
27
28 information about unit processes that constitute the incipient plasticity and inelastic response of the
29
30 material. MD simulations of nanoindentation of energetic molecular crystals are limited. Chen et al.40
31
32 studied nanoindentation on the (100) face of α-RDX using MD with a classical reactive force field. They
33
34
35 predicted significant heating near the indenter and limited pile-up. MD simulations of nanoindentation
36
37 and scratch testing of the (210) and (001) faces of α-RDX, using a classical non-reactive force field,
38
39 predicted anisotropy in reduced elastic modulus and dynamic coefficient of friction, and isotropic values
40
41 for hardness.50
42
43
44 The goal of the present study is to characterize the thermo-mechanical response and incipient
45
46 stages of inelastic/plastic deformation in TATB using MD simulations of nanoindentation. Simulations
47
48
were performed at 77 K for indentation on the three principal crystallographic planes of TATB—(100),
49
50
51 (010), and (001)—defined by the normal vectors b × c, c × a, and a × b, respectively, where a, b, and c
52
53 denote the lattice vectors of the triclinic unit cell. The general choice to study cryogenic conditions is
54
55 motivated by the low stacking-fault energies in the system combined with our goal of isolating the
56
57 anisotropic plasticity mechanisms. The specific choice to study 77 K – the normal boiling point of
58
59
60 5
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 61

1
2
3 nitrogen – is the relative ease with which such conditions might be achieved in the laboratory
4
5
6 should complementary experiments be undertaken.
7
8
9
10
11
12 2. Methods
13
14 All MD simulations were performed using the LAMMPS software53 in conjunction with the non-
15
16 polarizable, flexible-molecule force field for TATB originally developed by Bedrov et al.11 and
17
18 subsequently modified to include covalent bond stretching8 and intramolecular O-H repulsion.54 The force
19
20 field, details of which are available in refs. 8, 11, 54, and 55, has been used to predict TATB temperature-
21
22
and pressure-dependent elastic coefficients,11,19 anisotropic crystal thermal conductivity,8-10 surface-
23
24
25 initiated melting,54 and generalized stacking-fault energies.19 A cutoff distance of 11 Å was used for
26
27 repulsion, dispersion, and short-range Coulomb interactions. Long-range Coulomb interactions were
28
29 calculated using the Particle-Particle Particle-Mesh (PPPM)56 k-space solver in LAMMPS with the
30
31 relative RMS error in per-atom forces set to 10-6. Simulations of indentation on the (100), (010), and
32
33 (001) planes at T = 77 K were performed in the isochoric-isothermal (NVT) ensemble using the Nosé-
34
35
Hoover style57 equations of motion derived by Shinoda et al.58 The thermostat coupling parameter was
36
37
38 100.0 fs. Nanoindentation simulations in the isochoric-isoenergetic (NVE) ensemble were performed for
39
40 indentation on the (100) and (001) planes, at a total energy initially corresponding to 77 K, to investigate
41
42 localized heating under the indenter. Trajectory integration was performed using a velocity-Verlet
43
44 integrator59 with a time step of 0.25 fs and 0.1 fs for simulations in the NVT and NVE ensemble,
45
46 respectively. Atomic coordinates, velocities, and time-averaged per-atom stresses were saved every 250 fs
47
48
for post-processing. The 3D vector of forces acting on the indenter was saved every 12.5 fs. Illustrations
49
50
51 of the crystal structure and atomic configurations were created using the Visual Molecular Dynamics
52
53 (VMD)60 and the Open Visualization Tool (OVITO)61 software packages.
54
55
56
57
58
59
60 6
ACS Paragon Plus Environment
Page 7 of 61 The Journal of Physical Chemistry

1
2
3 Initial simulation cells were created using lattice parameters predicted by the force field at T = 77
4
5
6
K and P = 1 atm (a = 8.9776 Å, b = 8.9982 Å, c = 6.6537 Å, α = 103.9237°, β = 95.3833°, γ = 119.8670°). A 3D
7
8 periodic 26a×26b×35c simulation supercell consisting of 47,320 TATB molecules (1,135,680 atoms) was
9
10 obtained by replicating the triclinic unit cell by factors of 26, 26, and 35 along a, b, and c, respectively.
11
12 (An assessment of this simulation supercell size is provided in the Supporting Information.) Next, the
13
14 length of the simulation cell was extended in the a, b, or c direction for indentations perpendicular to the
15
16 (100), (010), or (001) crystallographic planes, respectively. The length of the vacant region was 170.57 Å,
17
18
19
179.96 Å, and 153.03 Å for the (100), (010), and (001) cases, respectively. This simulation cell was
20
21 equilibrated in the NVT ensemble for 25 ps at T = 77 K, with 3D periodic boundary conditions, to obtain
22
23 a thermalized, infinite half-space of TATB. Following equilibration, a single-unit-cell-thick region of
24
25 TATB molecules (1a×26b×35c, 26a×1b×35c, and 26a×26b×1c for indentations on the (100), (010), and
26
27 (001) planes, respectively) at the “bottom” of the respective simulation cell was held fixed to prevent
28
29 rigid-body motion of the cell during indentation.
30
31
32 The indenter was modeled as a purely repulsive, rigid, spherical indenter62 as implemented in
33
34
35 LAMMPS. This implementation has been used in multiple MD studies of nanoindentation62-66 and
36
37 circumvents the computational overhead associated with an explicit atomic representation of the indenter.
38
39 In addition, stress concentrations are minimal with a spherical indenter and this results in a distinct elastic
40
41 response at small indentation depths. In this implementation the indenter exerts a force F given by:
42
43
44 (1)
45
F(r) = − K(R − r)2 if r ≤ R
46 =0 if r > R .
47
48
49 Here r is the distance from the center of the indenter to a given atom, K is the force constant, and
50
51 R is the radius of the indenter. The radius of and force constant for the indenter were chosen to be 60 Å
52
53
and 77.4 kcal mol-1 Å-3 (3.35 eV Å-3), respectively. This force constant value corresponds to an indenter
54
55
56 elastic modulus of ≈539 GPa and is (post facto) approximately 20 times larger than the highest value for
57
58
59
60 7
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 61

1
2
3 indentation modulus for the planes considered in this study (see Table 1); it is well within the range, 1-10
4
5
6
eV Å-3, used in studies62-65 of atomic crystals with indentation moduli much larger than for TATB.
7
8 Displacement-controlled nanoindentation simulations were performed in a quasi-static manner by
9
10 incrementally advancing the indenter in steps of 1.0 Å with 2.5 ps of equilibration at each indentation
11
12 step. This results in an effective indentation velocity of 40 m s-1, which is approximately 1.2% of the
13
14 sound speed in TATB (3.27 km s-1, calculated using the Hill-average67 bulk modulus of 21 GPa19).
15
16
17 A schematic of the simulation setup is shown in Figure 2. The indenter is incrementally advanced
18
19
in directions normal to the (100), (010), and (001) planes given by b × c, c × a, and a × b, respectively.
20
21
22 The indentation direction is conveniently aligned with the z-axis of the Cartesian coordinate system (see
23
24 Figure 2) only in the case of the (001) plane. For indentations on (100) and (010) planes, appropriate
25
26 transformations of the simulation data were applied for post-processing, so that the indentation axis was
27
28 aligned with the z-axis of the Cartesian coordinate system. Thus, the various Cartesian-frame-specific
29
30 expressions below apply directly for all three cases.
31
32
33
34 Force-displacement data were calculated from the raw MD data using F MD = f n̂ , where FMD
35
36 is the force acting normal to the surface of indentation, f is the 3D vector of forces acting on the indenter
37
38 (obtained from LAMMPS output and averaged over the final 250 fs of 2.5 ps equilibration at each value
39
40
41 of indenter displacement), and n̂ is the unit vector normal to the indentation plane. The force-
42
43 displacement curves obtained from the simulations were compared to predictions of the analytical
44
45 solution based on Hertzian theory for the indentation of an anisotropic half-space with a frictionless, rigid
46
47 parabola of revolution,68–72 which is a good approximation for spherical indentation.65,72 The Hertzian
48
49 prediction for the force F at a given indenter displacement δ is:
50
51
52 (2)
4 1 3
53 F(δ ) = MR 2 δ 2 .
54 3
55
56
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 61 The Journal of Physical Chemistry

1
2
3 Calculation of the indentation modulus M in eq. (2) requires evaluation of the Barnett-Lothe tensor L and
4
5
6
the contact area between the indenter and the half-space. For the analytical solution, a circular contact
7
8 area is assumed (strictly true only when there is three- or four-fold rotational symmetry perpendicular to
9
10 the indentation plane68,71,72), and M is calculated using:
11
12
13 1 2π (3)
14 M=
4π ∫
0
L−133 (ω ) d ω .
15
16
17
18 The Barnett-Lothe tensor L in eq. (3) is calculated from the elastic stiffness tensor C, using the
19
20 fundamental elasticity matrix,73 in the right-handed Cartesian coordinate system (CS) x′-y′-z obtained by
21
22 incremental rotations of the x-y-z CS about the z-axis by an angle dω ; see Figure 2. Detailed steps for
23
24 evaluating eq. (3) can be found in refs. 65,71–73 and are also given in the Appendix.
25
26
27
28 The contact pressure P is calculated as P(δ ) = F(δ ) A(δ ) , where A is the projected contact
29
30 area calculated from the coordinates of the atoms that are in contact with the indenter (i.e., atoms that are
31
32 at a distance r = R from the center of the indenter). The coordinates of the atoms in contact with the
33
34 indenter are projected on to the (initial) indentation plane (x-y plane) and the contact area of the resulting
35
36
connected region is calculated using:64,74
37
38
39
40 π (4)
A= (xmax − xmin )( ymax − ymin ) .
41 4
42
43
44 Here, the subscripts ‘max’ and ‘min’ denote the maximum and minimum values of the x and y coordinates
45
46
of the atoms (after the projection onto the initial indentation plane) in contact with the indenter.
47
48
49
Thermal response during nanoindentation was investigated using NVE simulations with the same
50
51
52 simulation parameters described above. (A comparison of the force-displacement curves obtained for
53
54 indentation on the (001) plane using NVE and NVT simulations is provided in the Supporting
55
56 Information.) Components of the locally averaged stress tensor were calculated by performing a
57
58
59
60 9
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 61

1
2
3 non-weighted spatial average of the components of the time-averaged, per-atom stress tensor over a
4
5
6
sphere of radius rcut using
7
8
N ineigh N iat (5)
9
10 (σ ) =4
mn i
1
π rcut3
∑ (σ
j=1
per−atom
mn ),
11 j
3
12
13
14 where m, n =x, y, z; Nineigh is the number of neighbor molecular centers of mass (COMs) within a distance
15
16
rcut =11 Å of the COM of the ith molecule; Niat is the number of atoms in a molecule (Niat = 24 for TATB);
17
18
19 and σ mn
per−atom
are the components of the time-averaged, per-atom stress tensor obtained from LAMMPS
20
21
(averaged over 250 fs). Spatial distributions of kinetic energies, expressed in units of temperature, were
22
23
24 obtained by partitioning the molecular and atomic velocities to yield translational and ro-vibrational
25
26 components using the procedure originally developed by Strachan and Holian75 and implemented in refs.
27
28 76 and 77. In this scheme, a spatial average is used to define the local velocity of the molecule, v COM ,
29 i
30
31 which is calculated as:
32
33
34
35 ∑ w(r )v ij
COM
j
(6)
36 v COM = j
.
37
i
∑ w(r ) ij
38 j

39
40
41 Here, vjCOM represents the COM velocity vector of the jth neighbor of the ith molecule. Following previous
42
43 studies,75–77 the weighting function, w(rij), is chosen to be:
44
45
46 (7)
f (r)4
47 w(r) = if r ≤ χ
48 1+ f (r)4
49
50
w(r) = 0 if r > χ .
51
52
53 In eq. (7), f(r) = (r – χ)/σ, with χ = 3b (b is the lattice parameter in the b direction) and σ = b/10. The
54
55 inter
intermolecular (relative translational) temperature of the ith molecule, T , is defined using:
56 i
57
58
59
60 10
ACS Paragon Plus Environment
Page 11 of 61 The Journal of Physical Chemistry

1
2
3 (8)
1
4
5 3 inter

2 j
M j w(rij ) | v COM
j
− v COM
i
|2
k T = .
6
7
2 B i
∑ w(rij )
j
8
9
10
11 Here, Mj is the total mass of the molecule and kB is the Boltzmann constant. The intramolecular (ro-
12
13 vibrational) temperature of the ith molecule, Tiintra, is calculated using:
14
15
16 at
(9)
(3N iat − 3) N
1 i
17 k BTi = ∑ m j (v j − viCOM )2 .
intra

18 2 2 j=1
19
20
21 intra
22 The average intramolecular temperature in the neighborhood of the ith molecule, T , is obtained as:
i
23
24
25
26 intra
∑ w(r )T ij j
intra (10)
27 T = j
.
28
i
∑ w(r ) ij
29 j
30
31
32 total
33
The average total temperature, T , was calculated using a non-weighted spatial average over a
i
34
35 distance of χ = 3b using
36
37
38 N ineigh N iat (11)
3N ineigh N iat 1

total
39 kB T = m j v 2j .
40 2 i 2 j=1
41
42
43
It is important to understand that these ‘temperatures’ are really kinetic energies, rescaled to temperature
44
45
46 units (i.e., kinetic temperatures). In particular, we did not verify that the distributions of particle kinetic
47
48 energies used to calculate the various quantities correspond to the Maxwell-Boltzmann distribution.
49
50
51
52
53
54 3. Results and Discussion
55
56
57 3.1. Nanoindentation on the (001) plane
58
59
60 11
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 61

1
2
3 Results for nanoindentation on the (001) plane (also referred to as the basal plane) are discussed in this
4
5
6
subsection. Figure 3 shows, as functions of the indenter displacement, the force on the indenter in the
7
8 direction perpendicular to the indentation plane (hereafter referred to as the force-displacement curve,
9
10 panel a), the contact area (panel b), and the contact pressure (panel c), for nanoindentation on the basal
11
12 plane. The force on the indenter increases upon contact with the indentation plane and closely follows the
13
14 prediction from Hertzian theory (eqs. (2) and (3)), up to a displacement (δ) of 8 Å, indicating a distinct
15
16 elastic response. This can also be inferred from the agreement between MD predictions of contact area
17
18
19
and contact pressure, for δ ≤ 8 Å, and the corresponding predictions from Hertzian theory. The MD
20
21 prediction for the force-displacement curve steepens (stiffens) for δ > 8 Å then exhibits a load-drop at δ >
22
23 13 Å that signals the onset of inelastic deformation. This stiffening can also be observed in the contact
24
25 pressure (Figure 3c) for δ > 8 Å. Interestingly, the drop in the contact pressure occurs at δ > 10 Å (P >
26
27 2.3±0.2 GPa), indicating that the onset of inelastic deformation actually begins at a smaller value of
28
29 indenter displacement than predicted from the force-displacement curve. Figure 4a shows the distribution
30
31
32 of maximum shear stress, τmax, in the simulation cell at δ = 9 Å, (i.e, just before the drop in the contact
33
34 1
35 pressure). The maximum shear stress is calculated as τ max = σ − σ 3 , where σ1 > σ2 > σ3 are the
2 1
36
37
38 eigenvalues of the locally averaged per-atom stress tensor. As expected from Hertzian theory, the largest
39
40 value of the maximum shear stress occurs below the surface and on the indentation axis (the –z-axis
41
42 shown in Figure 4a). Figure 4b shows τmax as a function of normalized depth (|z|/rc, where rc is the contact
43
44 radius obtained from the Hertzian prediction for the contact area) for four values of the indentation
45
46 displacement δ = 7, 8, 9, and 10 Å. It can be observed from Figure 4b that the maximum shear stress
47
48
under the indenter increases with increasing δ and reaches a maximum of 1.45±0.13 GPa at δ = 9 Å (error
49
50
51 bars are not shown in Figure 4b for clarity). The drop in τmax at δ = 10 Å indicates the onset of inelastic
52
53 deformation and is consistent with the prediction from the drop in contact pressure. Note that the
54
55 maximum of τmax occurs at |z|/rc ≈ 0.43, which is close to the value of 0.4878 predicted for isotropic
56
57 materials with Poisson’s ratio ν = 0.3.
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 61 The Journal of Physical Chemistry

1
2
3 The nature of this inelastic deformation was investigated by calculating displacements between
4
5
6
nearest-neighbor molecular centers of mass, λij (δ) = rij(0)-rij(δ), described in refs. 76, 77, and 79. Here, j
7
8 denotes (all) the nearest-neighbors of the ith molecular COM in the un-deformed configuration (δ = 0),
9
10 rij(δ) is the vector joining an i-j COM pair in the deformed configuration (δ ≠ 0), and rij(0) is the vector
11
12 joining the same i-j pair in the un-deformed configuration. Figure 5a shows the projection of λij along the
13
14 c axis, plotted against the projection of λij along a, for the first 20 layers of molecules below the
15
16 indentation plane. It can be inferred from Figure 5a that the deformation is elastic up to δ = 9 Å. In
17
18
19
addition, it can be seen from Figure 5a that inelastic deformation occurs for δ ≥ 10 Å, with material slip
20
21 along the a direction in the basal plane. This agrees well with the previous predictions19 based on γ-
22
23 surfaces (see, e.g., refs. 80 and 81), adapted for flexible molecules,19,82,83 that slip in the a direction is
24
25 favored on the basal plane compared to the b and a-b directions.
26
27
28 The resolved shear stress in the a direction, τca, shown in Figure 5b, reaches an (absolute) critical
29
30 value of 0.42±0.06 GPa at δ = 9 Å before the onset of inelastic deformation. This resolved shear stress
31
32
can be compared to the theoretical shear strength, calculated using:
33
34
35
µ G (12)
36 τ TS = .
37 d 2π
38
39
40 Here, µ is the intermolecular spacing in the slip direction, d is the interplanar spacing for the basal plane,
41
42 and G is the shear modulus in the appropriate crystallographic direction. The values for µ and d were
43
44
45 calculated using the (77 K, 1 atm) lattice parameters. To account for the compression of the lattice under
46
47 the indenter, G was taken to be the value of elastic coefficient C55 at pressure 1.16 GPa. (The pressure of
48
49 1.16 GPa in the vicinity of the indenter (|z|=0) was calculated using the per-atom stress tensor and C55 was
50
51 obtained using linear interpolation between pressure-dependent elastic coefficients available in the
52
53 literature.19) The value of τTS for the a direction using these parameters is 0.53 GPa. Comparing the value
54
55 of τTS to the critical value of the resolved shear stress in the a direction, 0.42±0.06 GPa, it can be inferred
56
57
58 that the onset of inelastic deformation occurs close to the theoretical shear strength of the crystal.
59
60 13
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 61

1
2
3 Snapshots of the atomic configuration at selected values of δ are shown in Figure 6 to illustrate
4
5
6
the stages of inelastic deformation on the basal plane. The atoms in Figures 6a are colored according to
7
8 atomic Mises strain84,85 (εMises, calculated using OVITO). For δ < 12 Å the inelastic deformation is
9
10 confined to basal planes below the indentation surface (not shown). This is followed by delamination of
11
12 the basal planes at the surface at δ = 12 Å (see Figure 6a). As δ increases further, material slip occurs on
13
14 the basal planes resulting in irreversible kinking and delamination of the basal planes (see Figure 6b).
15
16 Irreversible kinking and delamination have been observed during spherical nanoindentation of layered
17
18
19
crystals such as graphite86 and MAX phases87 (ternary carbides with the formula Mn+1AXn, where n = 1–3,
20
21 M is an early transition metal, A is an A-group element, and X is C and/or N). However, the formation of
22
23 reversible incipient kink-bands88,89 observed in these materials was not identified in the present study.
24
25 Previous studies86,90 have shown that reversible incipient kink-bands form at resolved shear stress values
26
27 of approximately G/30, which is much lower than the values at which inelastic deformation is predicted in
28
29 our simulations (≈ τTS), and this suggests that such a mechanism is not active for TATB under the
30
31
32 conditions investigated in this study.
33
34
35 With increase in indentation depth, the slipped material flows along the basal planes to the
36
37 surface and results in pile-up around the indenter (see Figure 6c). This material pile-up persists after
38
39 complete unloading, as can be seen from the surface-mesh construction91 of the indentation imprint
40
41 (obtained using OVITO) shown in Figure 6d. The indentation imprint and the pile-up are both circular,
42
43 which indicates that material flow on the basal plane is not limited to the a direction; that is, during
44
45
nanoindentation, the resolved shear stresses on the basal plane become large enough to result in material
46
47
48 flow in other directions. This agrees qualitatively with predictions using flexible molecule γ-surfaces for
49
50 TATB19 that the barriers to slip (unstable stacking-fault energies) are small along the a, a-b, and b
51
52 directions in the basal plane (among which the a direction has the lowest barrier). Note that the contact
53
54 pressure remains approximately constant (see Figure 3c) in the regime of indenter displacements, 15 Å <
55
56 δ < 24 Å, for which material flow occurs and leads to pile-up. An upper bound of the nanoindentation
57
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 61 The Journal of Physical Chemistry

1
2
3 (Meyer’s) hardness of TATB, obtained by averaging the contact pressure in this regime, is 1.02±0.09
4
5
6
GPa. (Nanoindentation hardness is inversely proportional to indentation depth; see, e.g., ref. 92.) While
7
8 there is no experimental nanoindentation hardness value for TATB available for comparison, values for
9
10 other relevant polyatomic molecular crystals are given in Table 2. It can be seen that the nanoindentation
11
12 hardness of TATB is higher compared to the experimental values for the energetic molecular crystals α-
13
14 RDX42 and LLM-10551 but is similar to that for α-RDX obtained (at 300 K) from MD simulations.50 The
15
16 predicted nanoindentation hardness of TATB is similar to the experimental value reported for the
17
18
19
energetic molecular crystal β-HMX in ref. 43 but higher than the value reported in ref. 51. The
20
21 nanoindentation hardness of TATB is similar to experimental values for single-crystal acetaminophen93 (a
22
23 widely used analgesic) but smaller than that for sucrose41 (a model material for pharmaceuticals and
24
25 explosives).
26
27
28 Thermal response during nanoindentation was investigated using the same procedure outlined in
29
30 inter
31 Section II but with NVE rather than NVT trajectory integration. Figures 7a-d show the averaged T ,
i
32
33 intra total
34 T i
, and T
i
, under the indenter, as functions of the depth along the axis of indentation (–z-axis,
35
36
37 see Figure 4a) at different times during the equilibration at δ = 7 Å (which was chosen as a representative
38
39 value of δ in the elastic region of the force-displacement curve). The averaging was performed over the
40
41 cutoff radius (11 Å). Initially, the kinetic energy due to indenter displacement is deposited in the
42
43 translational degrees of freedom of the lattice (see Figure 7a). This results in a large overshoot in the
44
45 intermolecular temperature under the indenter (Figure 7a). This excitation propagates into the indented
46
47
48
half-space, as can be seen from the temperature profiles at 750 fs and 1250 fs (Figures 7b and 7c,
49
50 respectively). At 1250 fs the intermolecular, intramolecualr, and total temperatures under the indenter
51
52 begin to converge to the same value, indicating the approach toward local thermal equilibrium. Indeed,
53
54 these temperatures under the indenter are nearly indistinguishable by the end of the equilibration (2500 fs)
55
56 and this suggests that local thermal equilibrium is approached on that time scale. MD studies of energy
57
58
59
60 15
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 61

1
2
3 exchange between intermolecular and intramolecular degrees of freedom in TATB crystals at atmospheric
4
5
6
pressure have predicted relaxation times of approximately 10 ps.55 The faster equilibration between
7
8 intermolecular and intramolecular degrees of freedom predicted during nanoindentation can tentatively be
9
10 attributed to a combination of the high pressure (i.e., ≈0.5±0.1 GPa, calculated as , see the
11
12
13 inset to Figure 7d) and the stress gradients that exist under the indenter. Similar analysis at several values
14
15 of δ indicates that the total temperature increases to 474±34 K in the vicinity of the indenter (|z|=0) prior
16
17
18
to the onset of inelastic deformation.
19
20
21 3.2. Nanoindentation on the (100) and (010) planes
22
23
24 Results for nanoindentation on the (100) and (010) planes (non-basal planes) are discussed in this
25
26 subsection. Figure 8 shows, as functions of the indenter displacement for the (100) and (010)
27
28 indentations, MD results and Hertzian theory predictions for the force on the indenter in the direction
29
30 perpendicular to the respective indentation plane (panel a), the contact area (panel b), and the contact
31
32 pressure (panel c) obtained using NVT simulations. A softer elastic response for indentation on (010)
33
34 compared to (100) is predicted both by the MD simulations and Hertzian theory. Interestingly, the initial
35
36
37 parts of the force-displacement curve and the contact pressure obtained from MD simulations do not
38
39 follow the predictions from Hertzian theory; the MD simulations predict a more compliant response
40
41 compared to the Hertzian theory. This is different from the prediction for the basal plane, for which the
42
43 MD predictions and Hertzian theory were in close agreement. Note, however, that the MD result for the
44
45 contact area agrees well with the Hertzian prediction. From this we conclude that the anomalous, non-
46
47
Hertzian responses for (100) and (010) are due to an effective “softening” of the elastic-half space during
48
49
50 indentation and not due to imperfect contact between the sample and indenter. This softening is attributed
51
52 to the elastic bending of the molecular layers during nanoindentation. This is clearly visible in Figures 9a
53
54 and 9b, which show the atomic configurations at maximum indenter displacement, δ = 18 Å, for the (100)
55
56 and (010) indentations, respectively. The simulations were not continued for δ > 18 Å, as the strain fields
57
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 61 The Journal of Physical Chemistry

1
2
3 in the simulation cell begin to interact across the periodic boundaries for the system size used in this
4
5
6
study. Bending of the molecular layers does not persist upon complete unloading (Figures 9c and 9d for
7
8 the (100) and (010) indentations, respectively), although some residual deformation is evident (albeit with
9
10 noticeably less pile-up compared to the result shown in Figure 6 for nanoindentation on the basal plane).
11
12
13 Analysis of the thermal response was performed only for the (100) plane, as a representative case
14
15 inter intra total
16 for the non-basal planes. Figures 10a-d show the averaged T , T , and T under the
i i i
17
18 indenter, obtained from NVE simulations, as functions of the depth along the axis of indentation on the
19
20
(100) plane at four different times during the equilibration at δ = 7 Å. As predicted for indentation on the
21
22
23 basal plane, the kinetic energy due to the displacement of the indenter is initially deposited in the
24
25 translational degrees of freedom of the lattice. Interestingly the intermolecular temperature after 250 fs, in
26
27 the vicinity of the indenter, is considerably lower for the (100) plane (581±40 K) compared to that for the
28
29 basal plane (1357±92 K). Comparing the temperature profiles at 750 fs (Figures 7b and 10b) one can infer
30
31 that in both cases the excitation resulting from the indenter displacement propagates into the respective
32
33
half-spaces. Similar to the prediction for the basal plane, for the indenter depth shown local thermal
34
35
36 equilibrium is approximately achieved by the end of the 2500 fs equilibration period. Comparison of
37
total
38 T in the vicinity of the indenter at multiple values of the indenter displacement for the (100) and
39 i
40
41 (001) planes (see Figure 11) reveals that the temperature rise in the vicinity of the indenter is lower for the
42
43 (100) plane compared to the basal plane. This indicates that a smaller amount of mechanical energy is
44
45 converted to heat during indentation on the (100) plane compared to the basal plane. The maximum local
46
47
48
temperatures reached in our simulations are much lower than the decomposition threshold temperature of
49
50 1800 K predicted94 from reactive MD simulations of crystalline TATB. Also, Ramos et al.42 made no
51
52 mention of reactivity in their nanoindentation study of the more sensitive explosive RDX. For these
53
54 reasons, we think neglect of reactivity in our simulations is a reasonable approximation.
55
56
57
58
59
60 17
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 61

1
2
3 A comparison of mechanical response during NVT simulations of nanoindentation on the (100)
4
5
6
plane with indenters of radius R = 30, 60, and 90 Å is given in Figure 12. The ordinate in Figure 12 is the
7
8 force F divided by R1/2; this serves to highlight the effect of the indenter radius on the results. The
9
10 response is non-Hertzian for all three values of R, but a clear trend exists, namely that MD predictions of
11
12 the normalized force-displacement curves using a larger indenter more closely approximate the Hertzian
13
14 prediction. This shows that the effective softening during nanoindentation on the non-basal planes
15
16 (attributed to the bending of molecular layers) is dependent on the indenter radius, and suggests that the
17
18
19
Hertzian response would possibly be recovered in the limit of a large indenter radius. However, MD
20
21 simulations using an indenter with R > 90 Å are beyond our current computational capabilities and the
22
23 results obtained using R = 60 Å are considered as representative.
24
25
26 4. Conclusions
27
28
29 Molecular dynamics simulations of displacement-controlled nanoindentations on the (100), (010),
30
31 and (001) planes of TATB were performed using a rigid, spherical indenter. The elastic part of the force-
32
33 displacement curve on the (001) plane (basal plane) is predicted accurately by the analytical solution
34
35
obtained using Hertzian theory for the indentation of an anisotropic half-space with a rigid, frictionless
36
37
38 parabola of revolution. Inelastic deformation during nanoindentation of the basal plane results in kinking
39
40 and delamination of the basal planes and leads to significant pile-up. An upper bound of 1.02±0.09 GPa is
41
42 predicted for nanoindentation hardness on the basal plane of TATB. A non-Hertzian response is predicted
43
44 for nanoindentation on the (100) and (010) planes (non-basal planes) and this is attributed to elastic
45
46 bending of the molecular layers during indentation, which results in an effective softening of the elastic
47
48
half-space. Pile-up during nanoindentation on the non-basal planes is less significant compared to that on
49
50
51 the basal plane. Analysis of spatially averaged kinetic energies, obtained by partitioning the molecular
52
53 velocity into relative translational and ro-vibrational components, predicts fast (≈ 2.5 ps) equilibration
54
55 between the translational and ro-vibrational kinetic energies, which is tentatively attributed to the stress
56
57 state under the indenter. Significant heating occurs during nanoindentation on the basal and non-basal
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 61 The Journal of Physical Chemistry

1
2
3 planes, with a larger amount of heating predicted for the basal plane for the same value of indenter
4
5
6
displacement.
7
8
9 Acknowledgments
10
11
12 The U.S. Air Force Office of Scientific Research supported this research under Grant No. FA9550-14-1-
13
14 0091. The authors thank Matt Kroonblawd for useful discussions on thermal transport and vibrational
15
16 relaxation in TATB.
17
18
19
20
21
22 Supporting Information
23
24
25 Comparison of the nanoindentation response predicted by NVE and NVT simulations and assessment of
26
27 the simulation supercell size are available via the
28
29
30 Internet at http://pubs.acs.org.
31
32
33
34
35
36 APPENDIX: Calculation of the Barnett-Lothe tensor and indentation modulus
37
38
39 Let C denote the fourth-rank stiffness tensor given in the right-handed Cartesian coordinate
40
41
system x-y-z and let C′ denote the same in the rotated coordinate system x′-y′-z, obtained by a rotation
42
43
44 about z by an angle dω. In the Stroh formalism65,73 a general solution for the displacement and stress field
45
46 can be constructed by solving for the six eigenvalues pm (three complex conjugate pairs for materials with
47
48 physically admissible elastic constants) and eigenvectors am of the eigenvalue problem given as:
49
50
51
Q + p(R + R T ) + p 2 T  a = 0. (A1)
52
53
54
55 Here, the second-rank (3×3) tensors Q, R, and T are calculated as:
56
57
58
59
60 19
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 61

1
2
3
4 Q jk = C ′j1k1; R jk = C ′j1k 2 ; T jk = C ′j2k 2 . (A2)
5
6
7
The vectors  a ,b  , where  R + p T  a = b , corresponding to an eigenvalue pm (with positive
m m T m m m
8
9
10 imaginary part from the complex conjugate pairs m = 1, 2, 3) are the right-eigenvectors of the
11
12 fundamental elasticity matrix,73 N, calculated as:
13
14
15
16  N N 
N= ; N 1 = −T −1R T ; N 3 = RT −1R T − Q.
1 2
17 N 2 = T _1 ; (A3)
18  N 3 N1
T

 
19
20
21
The Barnett-Lothe tensor, L, can be calculated as:
22
23
24 L = −2iBBT ; B ≡  b1 b 2 b 3  . (A4)
25  
26
27
28 Here, the vectors a and b need to be normalized so that:
29
30  am 
31  bm a m   m  = 1. (A5)
32   b 
33
 
34
35
36 Equation (A4) is evaluated for every incremental rotation dω, ω ∈ 0,2π  , and the indentation modulus
37
38
39
is calculated by evaluating the integral in eq. (2).
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 TABLES
9
10
11 Table 1: Indentation modulus of the crystallographic planes in TATB obtained using eq. (3).
12
13 Indentation plane Indentation modulus, M (GPa)
14
15 b × c or (100) 27.7
16
17
18 c × a or (010) 26.0
19
20 a × b or (001) 10.7
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 21
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 61

1
2
3
4
5
6 Table 2: Comparison of nanoindentation hardness of TATB to other polyatomic molecular
7
8 crystals. All values are obtained from experiments unless denoted using the asterisk symbol (*),
9
10 in which case they are obtained from MD simulations.
11
12
13
Material Crystal Symmetry Indentation Plane Hardness (GPa) Reference
14
15
16 TATBa Triclinic a × b or (001) 1.02±0.09* This work
17
18 (basal)
19
20 α-RDX Orthorhombic (210) 0.672±0.035 42
21
22 (210) 0.798±0.030 50
23
24 (210) 1.06* 50
25
26 (021) 0.681±0.033 42
27
28 (001) 0.615±0.035 42
29
30
(001) 1.05* 50
31
32
33 β-HMX Monoclinic (010) 1.13±0.045 43
34
35 (010) 0.65±0.09 51
36
37 LLM-105 Monoclinic (010) 0.73±0.10 51
38
39 Sucrose Monoclinic (001) 1.57±0.07 41
40
41 Acetaminophen Monoclinic (011) 0.875±0.029 93
42
43 a
Hardness values are not reported for the (100) and (010) planes for TATB, as nanoindentation on these
44
45
planes could only be performed to a depth δ = 18 Å. See Section 3.2 for details.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 22
ACS Paragon Plus Environment
Page 23 of 61 The Journal of Physical Chemistry

1
2
3
4
5
FIGURES
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 1. Structure and anisotropy of TATB. (a) A TATB molecule. Carbon is cyan, nitrogen is
27
28 blue, oxygen is red, and hydrogen is white. (b) A TATB unit cell with lattice vectors a, b, c and
29
30
31 lattice angles α, β, γ. (c) Layered crystal structure of TATB shown using a projection along b.
32
33 Dark gray and light blue spheres represent molecular centers of mass on alternating basal planes.
34
35
36 The prime symbol (′) on the axis labels in this and subsequent figures emphasizes that the lattice
37
38 vectors of the unit cell do not lie on the plane of the paper; this is not to be confused with the
39
40
notation a′ = b×c, b′ = c×a, and c′ = a×b used in refs. 8-10. (d) Polar plot of orientation-
41
42
43 dependent Young’s modulus of TATB; units on the color scale are GPa.
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 23
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure 2. Schematic of the simulation setup. The vectors a, b, and c define the lattice vectors of
23
24 the triclinic TATB unit cell and form a non-orthogonal right-handed basis. The right-handed
25
26
27 Cartesian coordinate system is represented as x-y-z, with the origin located at the point of contact
28
29 of the indenter with the indentation plane. In the chosen frame, a is aligned with x, b lies in the
30
31
x–y plane and makes an angle of γ with a, and c is in the +z half-space and makes angles of β
32
33
34 with a and α with b.
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 24
ACS Paragon Plus Environment
Page 25 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
Figure 3. Comparison between MD results and Hertzian theory for nanoindentation on the (001)
49
50 (basal) plane. (a) Force-displacement curve. The force on the indenter was obtained by averaging
51
52 over the final 250 fs of 2.5 ps of equilibration at every displacement. The error bars correspond
53
54
55 to one standard deviation of the mean. (b) Contact area calculated using eq. (3). (c) Contact
56
57
58
59
60 25
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 26 of 61

1
2
3
pressure under the indenter. The error bars are obtained by propagating the uncertainty in the
4
5
6 forces using standard error propagation methods.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 26
ACS Paragon Plus Environment
Page 27 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 4. Maximum shear stress (τmax) in the simulation cell for nanoindentation on (001): (a)
22
23 Distribution of τmax just before the onset of inelastic deformation, i.e., δ = 9 Å. Spheres represent
24
25
26
molecular center-of-mass positions; units on the color scale are GPa. (b) Maximum shear stress
27
28 along the indentation axis (–z-axis shown in (a)), for δ = 7, 8, 9, and 10 Å.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 27
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 Figure 5. Nature of incipient inelastic deformation during nanoindentation normal to the (001)
28
29
30
plane. (a) Scatter plots of the components along c and along a of slip vectors (λij), for molecules
31
32 in the first 20 layers under the indenter, for four different indentation depths. (b) Resolved shear
33
34 stress τca in the a direction on the (001) plane at δ = 9 Å. Spheres in (b) represent positions of the
35
36
37 centers of mass of molecules; units on the color scale are GPa.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 28
ACS Paragon Plus Environment
Page 29 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 6. Stages of inelastic deformation during nanoindentation normal to (001). The colors in
35
36
37 (a), (b), and (c) denote the atomic Mises strain. (a) Delamination of basal planes at the surface
38
39 (denoted by the arrow) at δ = 12 Å. Inelastic deformation is contained below the surface for δ <
40
41
12 Å. (b) Kinking and delamination of basal planes below the surface (denoted using arrows and
42
43
44 boxes) at δ = 15 Å. (c) Pile-up around the indenter at δ = 24 Å, resulting from material flow on
45
46 the basal planes to the indentation surface (denoted using arrows). (d) Circular indentation
47
48
49
imprint and pile-up (obtained using surface mesh construction, see text) after complete unloading
50
51 from δ = 30 Å.
52
53
54
55
56
57
58
59
60 29
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 30 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Figure 7. Spatially averaged intermolecular, intramolecular and total temperatures below the
36
37
38 indenter, as a function of the depth, at various instants during the equilibration period at δ = 7 Å
39
40 for nanoindentation on (001). (a) 250 fs (b) 750 fs (c) 1250 fs and (d) 2500 fs. Inset to (d) shows
41
42
43 spatially averaged pressure [ ], below the indenter, as a function of depth along –z
44
45
46 axis for δ = 7 Å. The error bars on all the panels correspond to one standard deviation of the
47
48 mean.
49
50
51
52
53
54
55
56
57
58
59
60 30
ACS Paragon Plus Environment
Page 31 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50 Figure 8. Comparison between MD results and Hertzian theory for nanoindentation on the (100) and
51
52 (010) (non-basal) planes. (a) Force-displacement curve. The force on the indenter was obtained by
53
54 averaging over the final 250 fs of 2.5 ps of equilibration at every displacement. The error bars correspond
55
56 to one standard deviation of the mean. (b) Contact area calculated using eq. (3). (c) Contact pressure
57
58
59
60 31
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 32 of 61

1
2
3 under the indenter. The error bars are obtained by propagating the uncertainty in the forces using standard
4
5
6
error propagation methods.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 32
ACS Paragon Plus Environment
Page 33 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 9. Atomic configurations for nanoindentation of the (100) and (010) planes. (a) (100) at
31
32
33 maximum indenter displacement δ = 18 Å. (b) (010) at maximum indenter displacement δ = 18
34
35 Å. (c) (100) completely unloaded from δ = 18 Å. (d) (010) completely unloaded from δ = 18 Å.
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 33
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 34 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 10. Spatially averaged intermolecular, intramolecular and total temperatures below the
35
36
37 indenter, as a function of the depth, at various instants during the equilibration period for δ = 7 Å
38
39 for nanoindentation on (100). (a) 250 fs (b) 750 fs (c) 1250 fs and (d) 2500 fs. The error bars on
40
41 all the panels correspond to one standard deviation of the mean.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 34
ACS Paragon Plus Environment
Page 35 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Figure 11. Comparison of locally averaged total temperature in the vicinity of the indenter for
21
22 nanoindentation on the (100) and (001) planes.
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 35
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 36 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 12. Comparison of the mechanical response during nanoindentation of the (100) plane
20
21 with indenters of radius R = 30, 60, and 90 Å. The force F in the ordinate is divided by R1/2 to
22
23 highlight the effect of the indenter radius on the results.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 36
ACS Paragon Plus Environment
Page 37 of 61 The Journal of Physical Chemistry

1
2
3 REFERENCES
4
5
(1) Cady, H. H.; Larson, A. C. The Crystal Structure of 1,3,5-Triamino-2,4,6-Trinitrobenzene. Acta
6
7
8 Crystallogr. 1965, 18 (3), 485–496.
9
10
11 (2) Filippini, G.; Gavezzotti, A. The Crystal Structure of 1, 3, 5-Triamino-2, 4, 6-Trinitrobenzene:
12
13 Centrosymmetric or Non-Centrosymmetric? Chem. Phys. Lett. 1994, 231, 86–92.
14
15
16 (3) Kolb, J. R.; Rizzo, H. F. Growth of 1, 3, 5- Triamino- 2, 4, 6- trinitrobenzene (TATB) I.
17
18
19 Anisotropic Thermal Expansion. Propellants, Explos. Pyrotech. 1979, 4 (1), 10–16.
20
21
22 (4) Voigt-Martin, I. G.; Li, G.; Yakimanski, A. A.; Wolff, J. J.; Gross, H. Use of Electron Diffraction
23
24 and High-Resolution Imaging To Explain Why the Non-Dipolar 1,3,5-Triamino-2,4,6-
25
26
Trinitrobenzene Displays Strong Powder Second Harmonic Generation Efficiency. J. Phys. Chem.
27
28
29 A 1997, 101 (39), 7265–7276.
30
31
32 (5) Davidson, A. J.; Dias, R. P.; Dattelbaum, D. M.; Yoo, C.-S. “Stubborn” Triaminotrinitrobenzene:
33
34 Unusually High Chemical Stability of a Molecular Solid to 150 GPa. J. Chem. Phys. 2011, 135
35
36 (17), 174507.
37
38
39 (6) Sun, J.; Kang, B.; Xue, C.; Liu, Y.; Xia, Y. Crystal State of 1, 3, 5-Triamino-2, 4, 6-
40
41 Trinitrobenzene (TATB) Undergoing Thermal Cycling Process. J. Energ. Mater. 2010, 28 (3), 37–
42
43 41.
44
45
46
(7) Taylor, D. E. Intermolecular Forces and Molecular Dynamics Simulation of 1,3,5-Triamino-2,4,6-
47
48
49 Trinitrobenzene (TATB) Using Symmetry Adapted Perturbation Theory. J. Phys. Chem. A 2013,
50
51 117 (16), 3507–3520.
52
53
54 (8) Kroonblawd, M. P.; Sewell, T. D. Theoretical Determination of Anisotropic Thermal Conductivity
55
56 for Crystalline 1,3,5-Triamino-2,4,6-Trinitrobenzene (TATB). J. Chem. Phys. 2013, 139 (7),
57
58
59
60 37
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 38 of 61

1
2
3 74503.
4
5
6 (9) Kroonblawd, M. P.; Sewell, T. D. Theoretical Determination of Anisotropic Thermal Conductivity
7
8
9 for Initially Defect-Free and Defective TATB Single Crystals. J. Chem. Phys. 2014, 141 (18),
10
11 184501.
12
13
14 (10) Kroonblawd, M. P.; Sewell, T. D. Predicted Anisotropic Thermal Conductivity for Crystalline
15
16 1,3,5-Triamino-2,4,6-Trinitobenzene (TATB): Temperature and Pressure Dependence and
17
18 Sensitivity to Intramolecular Force Field Terms. Propellants, Explos. Pyrotech. 2015, (in press);
19
20 DOI:10.1002/prep.201500247.
21
22
23 (11) Bedrov, D.; Borodin, O.; Smith, G. D.; Sewell, T. D.; Dattelbaum, D. M.; Stevens, L. L. A
24
25
Molecular Dynamics Simulation Study of Crystalline 1,3,5-Triamino-2,4,6-Trinitrobenzene as a
26
27
28 Function of Pressure and Temperature. J. Chem. Phys. 2009, 131 (22), 224703.
29
30
31 (12) Budzevich, M. M.; Landerville, A. C.; Conroy, M. W.; Lin, Y.; Oleynik, I. I.; White, C. T.
32
33 Hydrostatic and Uniaxial Compression Studies of 1,3,5-Triamino- 2,4,6-Trinitrobenzene Using
34
35 Density Functional Theory with van Der Waals Correction. J. Appl. Phys. 2010, 107 (11), 113524.
36
37
38 (13) Manaa, M. R.; Fried, L. E. Nearly Equivalent Inter- and Intramolecular Hydrogen Bonding in
39
40 1,3,5-Triamino-2,4,6-Trinitrobenzene at High Pressure. J. Phys. Chem. C 2012, 116 (3), 2116–
41
42 2122.
43
44
45 (14) Ojeda, O. U.; Çağın, T. Hydrogen Bonding and Molecular Rearrangement in 1,3,5-Triamino-
46
47
48
2,4,6-Trinitrobenzene under Compression. J. Phys. Chem. B 2011, 115 (42), 12085–12093.
49
50
51 (15) Valenzano, L.; Slough, W. J.; Perger, W. Accurate Prediction of Second-Order Elastic Constants
52
53 from First Principles: PETN and TATB. In AIP Conference Proceedings; 2012; Vol. 1191, pp
54
55 1191–1194.
56
57
58
59
60 38
ACS Paragon Plus Environment
Page 39 of 61 The Journal of Physical Chemistry

1
2
3 (16) Zhang, C. Investigation of the Slide of the Single Layer of the 1,3,5-Triamino-2,4,6-
4
5
6
Trinitrobenzene Crystal: Sliding Potential and Orientation. J. Phys. Chem. B 2007, 111 (51),
7
8 14295–14298.
9
10
11 (17) Zhang, C.; Wang, X.; Huang, H. Pi-Stacked Interactions in Explosive Crystals: Buffers against
12
13 External Mechanical Stimuli. J. Am. Chem. Soc. 2008, 130 (26), 8359–8365.
14
15
16 (18) Ting, T. C. T. On Anisotropic Elastic Materials for Which Young’s Modulus E(n) Is Independent
17
18 of N or the Shear Modulus G(n,m) Is Independent of N and M. J. Elast. 2005, 81 (3), 271–292.
19
20
21 (19) Mathew, N.; Sewell, T. D. Generalized Stacking Fault Energies in the Basal Plane of Triclinic
22
23 Molecular Crystal 1,3,5-Triamino-2,4,6-Trinitrobenzene (TATB). Philos. Mag. 2015, 95 (4), 424–
24
25
440.
26
27
28
(20) Phillips, D. S.; Schwarz, R. B.; Skidmore, C. B.; Hiskey, M. A.; Son, S. F. Some Observations on
29
30
31 the Structure of TATB. In AIP Conference Proceedings; AIP, 2000; Vol. 505, pp 707–710.
32
33
34 (21) Kennedy, J. E.; Lee, K.-Y.; Son, S. F.; Martin, E. S.; Asay, B. W.; Skidmore, C. B. Second-
35
36 Harmonic Generation and the Shock Sensitivity of TATB. In AIP Conference Proceedings; AIP,
37
38 2000; Vol. 505, pp 711–714.
39
40
41 (22) Hoffman, D. M.; Fontes, A. Density Distributions in TATB Prepared by Various Methods.
42
43 Propellants, Explos. Pyrotech. 2009, 94550, 15–23.
44
45
46 (23) Bowden, F. P.; Yoffe, A. D. Initiation and Growth of Explosion in Liquids and Solids, 1st ed.;
47
48 Cambridge University Press: Cambridge,UK, 1985.
49
50
51
(24) Field, J. E. Hot Spot Ignition Mechanisms for Explosives. Acc. Chem. Res. 1992, 25 (11), 489–
52
53
54 496.
55
56
57 (25) Cawkwell, M.; Sewell, T.; Zheng, L.; Thompson, D. Shock-Induced Shear Bands in an Energetic
58
59
60 39
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 40 of 61

1
2
3 Molecular Crystal: Application of Shock-Front Absorbing Boundary Conditions to Molecular
4
5
6
Dynamics Simulations. Phys. Rev. B 2008, 78 (1), 014107.
7
8
9 (26) Dattelbaum, D. M.; Sheffield, S. A.; Stahl, D. B.; Dattelbaum, A. M.; Elert, M.; Furnish, M. D.;
10
11 Anderson, W. W.; Proud, W. G.; Butler, W. T. Influence of Hot Spot Features on the Shock
12
13 Initiation of Heterogeneous Nitromethane. In AIP Conference Proceedings; 2009; Vol. 263, pp
14
15 263–266.
16
17
18 (27) An, Q.; Goddard, W. A.; Zybin, S. V; Jaramillo-Botero, A.; Zhou, T. Highly Shocked Polymer
19
20 Bonded Explosives at a Nonplanar Interface: Hot-Spot Formation Leading to Detonation. J. Phys.
21
22
Chem. C 2013, 117 (50), 26551–26561.
23
24
25
(28) Wood, M. A.; Cherukara, M. J.; Kober, E. M.; Strachan, A. Ultrafast Chemistry under
26
27
28 Nonequilibrium Conditions and the Shock to Deflagration Transition at the Nanoscale. J. Phys.
29
30 Chem. C 2015, 119 (38), 22008–22015.
31
32
33 (29) You, S.; Chen, M.-W.; Dlott, D. D.; Suslick, K. S. Ultrasonic Hammer Produces Hot Spots in
34
35 Solids. Nat. Commun. 2015, 6, 6581.
36
37
38 (30) Joshi, K. L.; Chaudhuri, S. Reactive Simulation of the Chemistry behind the Condensed-Phase
39
40 Ignition of RDX from Hot Spots. Phys. Chem. Chem. Phys. 2015, 17, 18790–18801.
41
42
43 (31) Dlott, D. D.; Fayer, M. D. Shocked Molecular Solids: Vibrational up Pumping, Defect Hot Spot
44
45 Formation, and the Onset of Chemistry. J. Chem. Phys. 1990, 92 (6), 3798.
46
47
48 (32) Dick, J. J.; Mulford, R. N.; Spencer, W. J.; Pettit, D. R.; Garcia, E.; Shaw, D. C. Shock Response
49
50
51 of Pentaerythritol Tetranitrate Single Crystals. J. Appl. Phys. 1991, 70 (7), 3572.
52
53
54 (33) Coffey, C. S. The Localization of Energy and Plastic Deformation in Crystalline Solids during
55
56 Shock or Impact. J. Appl. Phys. 1991, 70 (8), 4248–4254.
57
58
59
60 40
ACS Paragon Plus Environment
Page 41 of 61 The Journal of Physical Chemistry

1
2
3 (34) Armstrong, R. W.; Ammon, H. L.; Elban, W. L.; Tsai, D. H. Investigation of Hot Spot
4
5
6
Characteristics in Energetic Crystals. Thermochim. Acta 2002, 384 (1-2), 303–313.
7
8
9 (35) Armstrong, R. W.; Elban, W. L. Materials Science and Technology Aspects of Energetic
10
11 (explosive) Materials. Mater. Sci. Technol. 2006, 22 (4), 381–395.
12
13
14 (36) Armstrong, R. W. Dislocation Mechanics Aspects of Energetic Material Composites. Rev. Adv.
15
16 Mater. Sci 2009, 19, 13–40.
17
18
19 (37) Dick, J. J.; Ritchie, J. P. Molecular Mechanics Modeling of Shear and the Crystal Orientation
20
21 Dependence of the Elastic Precursor Shock Strength in Pentaerythritol Tetranitrate. J. Appl. Phys.
22
23 1994, 76 (5), 2726–2737.
24
25
26 (38) Dick, J. J. Anomalous Shock Initiation of Detonation in Pentaerythritol Tetranitrate Crystals. J.
27
28
Appl. Phys. 1997, 81, 601–612.
29
30
31
(39) Liao, X.; Wiedmann, T. S. Characterization of Pharmaceutical Solids by Scanning Probe
32
33
34 Microscopy. J. Pharm. Sci. 2004, 93 (9), 2250–2258.
35
36
37 (40) Chen, Y.-C.; Nomura, K.; Kalia, R. K.; Nakano, A.; Vashishta, P. Molecular Dynamics
38
39 Nanoindentation Simulation of an Energetic Material. Appl. Phys. Lett. 2008, 93 (17), 171908.
40
41
42 (41) Ramos, K.; Bahr, D. Mechanical Behavior Assessment of Sucrose Using Nanoindentation. J.
43
44 Mater. Res. 2007, 22 (7), 2037–2045.
45
46
47 (42) Ramos, K. J.; Hooks, D. E.; Bahr, D. F. Direct Observation of Plasticity and Quantitative
48
49 Hardness Measurements in Single Crystal Cyclotrimethylene Trinitramine by Nanoindentation.
50
51
Philos. Mag. 2009, 89 (27), 2381–2402.
52
53
54
(43) Li, M.; Tan, W. J.; Kang, B.; Xu, R. J.; Tang, W. The Elastic Modulus of β-HMX Crystals
55
56
57 Determined by Nanoindentation. Propellants, Explos. Pyrotech. 2010, 35, 379–383.
58
59
60 41
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 42 of 61

1
2
3 (44) Ramos, K. J.; Bahr, D. F.; Hooks, D. E. Defect and Surface Asperity Dependent Yield during
4
5
6
Contact Loading of an Organic Molecular Single Crystal. Philos. Mag. 2011, 91 (7-9), 1276–1285.
7
8
9 (45) Armstrong, R. W.; Elban, W. L. Macro- to Nano-Indentation Hardness Stress–Strain Aspects of
10
11 Crystal Elastic/Plastic/Cracking Behaviors. Exp. Mech. 2010, 50 (4), 545–552.
12
13
14 (46) Reddy, C. M.; Rama Krishna, G.; Ghosh, S. Mechanical Properties of Molecular Crystals—
15
16 applications to Crystal Engineering. Cryst. Eng. Comm. 2010, 12 (8), 2296.
17
18
19 (47) Jing, Y.; Zhang, Y.; Blendell, J.; Koslowski, M.; Carvajal, M. T. Nanoindentation Method to
20
21 Study Slip Planes in Molecular Crystals in a Systematic Manner. Cryst. Growth Des. 2011, 11
22
23 (12), 5260–5267.
24
25
26 (48) Mishra, M. K.; Ramamurty, U.; Desiraju, G. R. Solid Solution Hardening of Molecular Crystals:
27
28
Tautomeric Polymorphs of Omeprazole. J. Am. Chem. Soc. 2015, 137 (5), 1794–1797.
29
30
31
(49) Maughan, M. R.; Carvajal, M. T.; Bahr, D. F. Nanomechanical Testing Technique for Millimeter-
32
33
34 Sized and Smaller Molecular Crystals. Int. J. Pharm. 2015, 486 (1-2), 324–330.
35
36
37 (50) Weingarten, N. S.; Sausa, R. C. Nanomechanics of RDX Single Crystals by Force–Displacement
38
39 Measurements and Molecular Dynamics Simulations. J. Phys. Chem. A 2015, 119 (35), 9338–
40
41 9351.
42
43
44 (51) Kucheyev, S. O.; Gash, A. E.; Lorenz, T. Deformation and Fracture of LLM-105 Molecular
45
46 Crystals Studied by Nanoindentation. Mater. Res. Express 2014, 1 (2), 025036.
47
48
49 (52) Hudson, R. J.; Zioupos, P.; Gill, P. P. Investigating the Mechanical Properties of RDX Crystals
50
51
Using Nano-Indentation. Propellants, Explos. Pyrotech. 2012, 37 (2), 191–197.
52
53
54
(53) Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys.
55
56
57 1995, 117, 1–19; Also see http://lammps.sandia.gov/.
58
59
60 42
ACS Paragon Plus Environment
Page 43 of 61 The Journal of Physical Chemistry

1
2
3 (54) Mathew, N.; Sewell, T. D.; Thompson, D. L. Anisotropy in Surface-Initiated Melting of the
4
5
6
Triclinic Molecular Crystal 1,3,5-Triamino-2,4,6-Trinitrobenzene: A Molecular Dynamics Study.
7
8 J. Chem. Phys. 2015, 143 (9), 094706.
9
10
11 (55) Kroonblawd, M. P.; Sewell, T. D.; Maillet, J.-B. Characteristics of Energy Exchange between
12
13 Inter- and Intramolecular Degrees of Freedom in Crystalline 1,3,5-Triamino-2,4,6-Trinitrobenzene
14
15 (TATB) with Implications for Coarse-Grained Simulations of Shock Waves in Polyatomic
16
17 Molecular Crystals. J. Chem. Phys. 2016, 144 (6), 064501.
18
19
20 (56) Hockney, R. W.; Eastwood, J. W. Computer Simulation Using Particles, Paperback.; Taylor &
21
22
Francis: New York,NY, 1989.
23
24
25
(57) Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space Distributions. Phys. Rev. A 1985,
26
27
28 31 (3), 1695–1697.
29
30
31 (58) Shinoda, W.; Shiga, M.; Mikami, M. Rapid Estimation of Elastic Constants by Molecular
32
33 Dynamics Simulation under Constant Stress. Phys. Rev. B 2004, 69 (13), 134103.
34
35
36 (59) Tuckerman, M. E.; Alejandre, J.; López-Rendón, R.; Jochim, A. L.; Martyna, G. J. A Liouville-
37
38 Operator Derived Measure-Preserving Integrator for Molecular Dynamics Simulations in the
39
40 Isothermal–isobaric Ensemble. J. Phys. A. Math. Gen. 2006, 39 (19), 5629.
41
42
43 (60) Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996,
44
45 14 (1), 33–38;Also see http://www.ks.uiuc.edu/Research/vmd.
46
47
48 (61) Stukowski, A. Visualization and Analysis of Atomistic Simulation Data with OVITO–the Open
49
50
51 Visualization Tool. Model. Simul. Mater. Sci. Eng. 2009, 18 (1), 015012; Also see
52
53 http://www.ovito.org/.
54
55
56 (62) Kelchner, C.; Plimpton, S.; Hamilton, J. Dislocation Nucleation and Defect Structure during
57
58
59
60 43
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 44 of 61

1
2
3 Surface Indentation. Phys. Rev. B 1998, 58 (17), 85–88.
4
5
6 (63) Lilleodden, E. T.; Zimmerman, J. A.; Foiles, S. M.; Nix, W. D. Atomistic Simulations of Elastic
7
8
9 Deformation and Dislocation Nucleation during Nanoindentation. J. Mech. Phys. Solids 2003, 51
10
11 (5), 901–920.
12
13
14 (64) Ziegenhain, G.; Urbassek, H. M.; Hartmaier, A. Influence of Crystal Anisotropy on Elastic
15
16 Deformation and Onset of Plasticity in Nanoindentation: A Simulational Study. J. Appl. Phys.
17
18 2010, 107 (6), 061807.
19
20
21 (65) Zhu, T.; Li, J.; Van Vliet, K. J.; Ogata, S.; Yip, S.; Suresh, S. Predictive Modeling of
22
23 Nanoindentation-Induced Homogeneous Dislocation Nucleation in Copper. J. Mech. Phys. Solids
24
25
2004, 52 (3), 691–724.
26
27
28
(66) Remington, T. P.; Ruestes, C. J.; Bringa, E. M.; Remington, B. A.; Lu, C. H.; Kad, B.; Meyers, M.
29
30
31 A. Plastic Deformation in Nanoindentation of Tantalum: A New Mechanism for Prismatic Loop
32
33 Formation. Acta Mater. 2014, 78, 378–393.
34
35
36 (67) Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. Sect. A 1952, 65 (5),
37
38 349.
39
40
41 (68) Willis, J. R. Hertzian Contact of Anisotropic Bodies. J. Mech. Phys. Solids 1966, 14 (3), 163–176.
42
43
44 (69) Barnett, D. M.; Lothe, J. Line Force Loadings on Anisotropic Half-Spaces and Wedges. Phys.
45
46 Nor. 1996, 8, 163.
47
48
49 (70) Vlassak, J. J.; Nix, W. D. Indentation Modulus of Elastically Anisotropic Half Spaces. Philos.
50
51
Mag. A 1993, 67 (5), 1045–1056.
52
53
54
(71) Vlassak, J. J.; Nix, W. Measuring the Elastic Properties of Anisotropic Materials by Means of
55
56
57 Indentation Experiments. J. Mech. Phys. Solids 1994, 42 (8), 1223–1245.
58
59
60 44
ACS Paragon Plus Environment
Page 45 of 61 The Journal of Physical Chemistry

1
2
3 (72) Swadener, J. G.; Pharr, G. M. Indentation of Elastically Anisotropic Half-Spaces by Cones and
4
5
6
Parabolae of Revolution. Philos. Mag. A 2001, 81 (2), 447–466.
7
8
9 (73) Bower, A. F. Applied Mechanics of Solids; Taylor & Francis: Boca Raton, Florida, 2010.
10
11
12 (74) Gao, Y.; Ruestes, C. J.; Tramontina, D. R.; Urbassek, H. M. Comparative Simulation Study of the
13
14 Structure of the Plastic Zone Produced by Nanoindentation. J. Mech. Phys. Solids 2015, 75, 58–
15
16 75.
17
18
19 (75) Strachan, A.; Holian, B. L. Energy Exchange between Mesoparticles and Their Internal Degrees of
20
21 Freedom. Phys. Rev. Lett. 2005, 94 (1), 14301.
22
23
24 (76) Jaramillo, E.; Sewell, T. D.; Strachan, A. Atomic-Level View of Inelastic Deformation in a Shock
25
26 Loaded Molecular Crystal. Phys. Rev. B 2007, 76 (6), 064112.
27
28
29 (77) Eason, R. M.; Sewell, T. D. Shock-Induced Inelastic Deformation in Oriented Crystalline
30
31
32
Pentaerythritol Tetranitrate. J. Phys. Chem. C 2012, 116 (3), 2226–2239.
33
34
35
(78) Johnson, K. L. Contact Mechanics; Cambridge University Press: Cambridge, 1985.
36
37
38
(79) Zimmerman, J. A.; Kelchner, C. L.; Klein, P. A.; Hamilton, J. C.; Foiles, S. M. Surface Step
39
40 Effects on Nanoindentation. Phys. Rev. Lett. 2001, 87 (16), 165507.
41
42
43 (80) Hull, D.; Bacon, D. J. Introduction to Dislocations, 4th ed.; Butterworth-Heinemann: Oxford,UK,
44
45 2001.
46
47
48 (81) Duesbery, M. S.; Vitek, V. Plastic Anisotropy in B.c.c. Transition Metals. Acta Mater. 1998, 46
49
50 (5), 1481–1492.
51
52
53 (82) Cawkwell, M. J.; Ramos, K. J.; Hooks, D. E.; Sewell, T. D. Homogeneous Dislocation Nucleation
54
55 in Cyclotrimethylene Trinitramine under Shock Loading. J. Appl. Phys. 2010, 107 (6), 063512.
56
57
58
59
60 45
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 46 of 61

1
2
3 (83) Munday, L. B.; Solares, S. D.; Chung, P. W. Generalized Stacking Fault Energy Surfaces in the
4
5
6
Molecular Crystal RDX. Philos. Mag. 2012, 92 (December 2012), 37–41.
7
8
9 (84) Falk, M. L.; Langer, J. S. Dynamics of Viscoplastic Deformation in Amorphous Solids. Phys. Rev.
10
11 E 1998, 57 (6), 7192–7205.
12
13
14 (85) Shimizu, F.; Ogata, S.; Li, J. Theory of Shear Banding in Metallic Glasses and Molecular
15
16 Dynamics Calculations. Mater. Trans. 2007, 48 (11), 2923–2927.
17
18
19 (86) Barsoum, M. W.; Murugaiah, A.; Kalidindi, S. R.; Zhen, T.; Gogotsi, Y. Kink Bands, Nonlinear
20
21 Elasticity and Nanoindentations in Graphite. Carbon N. Y. 2004, 42 (8-9), 1435–1445.
22
23
24 (87) Molina-Aldareguiaa, J. M.; Emmerlich, J.; Palmquist, J.-P.; Jansson, U.; Hultman, L. Kink
25
26 Formation around Indents in Laminated Ti3SiC2 Thin Films Studied in the Nanoscale. Scr. Mater.
27
28
2003, 49 (2), 155–160.
29
30
31
(88) Barsoum, M. W.; Zhen, T.; Kalidindi, S. R.; Radovic, M.; Murugaiah, A. Fully Reversible,
32
33
34 Dislocation-Based Compressive Deformation of Ti3SiC2 to 1 GPa. Nat. Mater. 2003, 2 (2), 107–
35
36 111.
37
38
39 (89) Barsoum, M. W.; Murugaiah, A.; Kalidindi, S. R.; Zhen, T. Kinking Nonlinear Elastic Solids,
40
41 Nanoindentations, and Geology. Phys. Rev. Lett. 2004, 92 (25), 255508.
42
43
44 (90) Frank, F. C.; Stroh, A. N. On the Theory of Kinking. Proc. Phys. Soc. Sect. B 1952, 65 (10), 811–
45
46 821.
47
48
49 (91) Stukowski, A. Computational Analysis Methods in Atomistic Modeling of Crystals. JOM 2014, 66
50
51
(3), 399–407.
52
53
54
(92) Nix, W. D.; Gao, H. Indentation Size Effects in Crystalline Materials: A Law for Strain Gradient
55
56
57 Plasticity. J. Mech. Phys. Solids 1998, 46 (3), 411–425.
58
59
60 46
ACS Paragon Plus Environment
Page 47 of 61 The Journal of Physical Chemistry

1
2
3 (93) Shi, L.; Sun, C. C. Overcoming Poor Tabletability of Pharmaceutical Crystals by Surface
4
5
6
Modification. Pharm. Res. 2011, 28 (12), 3248–3255.
7
8
9 (94) Zhang, L.; Zybin, S. V; van Duin, A. C. T.; Dasgupta, S.; Goddard, W. A.; Kober, E. M. Carbon
10
11 Cluster Formation during Thermal Decomposition of Octahydro-1,3,5,7-Tetrazocine and 1,3,5-
12
13 Triamino-2,4,6-Trinitrobenzene High Explosives from ReaxFF Reactive Molecular Dynamics
14
15 Simulations. J. Phys. Chem. A 2009, 113 (40), 10619–10640.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Table of Contents (TOC) Image
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 47
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 48 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 48
ACS Paragon Plus Environment
Page 49 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Figure 1.
33 83x62mm (300 x 300 DPI)
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 50 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
Figure 3.
47
83x177mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 51 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 4.
20 166x58mm (300 x 300 DPI)
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 52 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Figure 6.
33 163x123mm (300 x 300 DPI)
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 53 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 9.
29 166x103mm (300 x 300 DPI)
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 54 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 10.
32 165x118mm (300 x 300 DPI)
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 55 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 11.
31
82x58mm (300 x 300 DPI)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 56 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 12.
31
82x58mm (300 x 300 DPI)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 57 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Table of Contents [TOC] Graphic
29 81x50mm (300 x 300 DPI)
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 58 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 11.
31
82x58mm (300 x 300 DPI)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 59 of 61 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 12.
31
82x58mm (300 x 300 DPI)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 60 of 61

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Table of Contents [TOC] Graphic
29 81x50mm (300 x 300 DPI)
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

S-ar putea să vă placă și