Mechanical Properties and Defect Sensitivity of Diamond

Compressive failure of a carbon nano-tesseract: Sci-Fi inspired materials and the strength of thanos. Steven W. Cranford. Extreme Mechanics Letters 20...
4 downloads 0 Views 8MB Size
Subscriber access provided by UNIV PRINCE EDWARD ISLAND

Communication

Mechanical Properties and Defect Sensitivity of Diamond Nanothreads Ruth Roman, Kenny Kwan, and Steven Cranford Nano Lett., Just Accepted Manuscript • DOI: 10.1021/nl5041012 • Publication Date (Web): 18 Feb 2015 Downloaded from http://pubs.acs.org on February 21, 2015

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.

Nano Letters 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 25

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

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Nano Letters

Mechanical Properties and Defect Sensitivity of Diamond Nanothreads Ruth E. Roman1, Kenny Kwan1, and Steven W. Cranford1* 1

Laboratory of Nanotechnology In Civil Engineering, Department of Civil and Environmental Engineering, Northeastern University, Boston, MA 02115 *Corresponding author: [email protected]

Abstract: One of the newest carbon allotropes synthesized are diamond nanothreads. Using molecular dynamics, we determine stiffness (850 GPa), strength (26.4 nN), extension (14.9%), and bending rigidity (.  ×  N-m2). The 1D nature of the nanothread results in a tenacity of .  ×  N-m/kg, exceeding nanotubes and graphene. As the thread consists of repeating Stone-Wales defects, through steered molecular dynamics (SMD), we explore the effect of defect density on the strength, stiffness and extension of the system.

21

Keywords: carbon allotropes; diamond nanothreads; mechanical properties; Stone-Wales

22

defects; molecular dynamics

23 24

The continuing focus of improvement and development of nanotechnology has unequivocally

25

made carbon allotropes one of the key subjects of material science research. Their inherent

26

molecular stability, in combination with promising electrical and thermal properties, spanning

27

different forms of allotropes,1-10 have enticed the material science community. Recently, a new

28

structural form of carbon was successfully synthesized from benzene precursors into a so-called

29

diamond nanothread.11 The core of the nanothread is a long, thin strand of carbon atoms arranged

30

in a diamond-like tetrahedral motif (a close-packed sp3-bonded carbon structure), the first

31

member of a new class of sp3 materials created under high-pressure solid-state reactions.

32

Uniquely, the presence of multiple covalent bonds across their cross-section removes them from

33

the realm of traditional polymers while the lack of a central hollow disqualifies the term

34

nanotube. Such threads represent a hybrid molecular structure, resembling a 3D molecular truss.

ACS Paragon Plus Environment

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 2 of 25 Roman et al.

35 36

The diamond thread is the result of controlled high-pressure conditions that force the reactions of

37

benzene “chains” into an semi-ordered one-dimensional (1D) saturated nanomaterial that is

38

formed by linked diamondoid-like structures (see Fig.1A; additional description in Supporting

39

Information). Based on the matching between experimental and modeling work,11 the nanothread

40

can be considered variation of a hydrogenated (3,0) nanotube12 with distributed Stone-Wales

41

transformation defects (C-C dimers rotated by 90ο). Note that, due to the 1:1 C/H stoichiometric

42

ratio, the thread can be thought of not as diamond (which is pure carbon), but rather a rolled

43

sheet of hydrogenated graphene (or graphene), which has pure sp3 bonding and the necessary

44

hydrogens. Due to the sp3 structure, the mechanical properties of such threads may be promising,

45

but are currently unknown. Understanding of the mechanical behavior including limit states such

46

as fracture stress and maximum strain is critical to explore the usability of such nanothreads as a

47

nanomaterial, and to extend its use within nanobundles and/or nanofabrics. To this end, here the

48

mechanical properties of the nanothread are explored considering an atomistically large model,

49

using full atomistic molecular dynamics (MD).

50 51

We apply a suite of in silico testing methods where full atomistic calculations of mechanical test

52

cases is implemented to derive a set of parameters to characterize the diamond nanothread,

53

similar to methods characterizing previous molecular systems.3-5, 13 We first confirm our model

54

is structurally similar to the synthesized nanothreads. The axial stiffness as well as the

55

strength/strain relationship are then investigated (via both quasi-static and dynamic steered

56

molecular dynamic (SMD) approaches), bending rigidity is defined (via energy minimization

57

molecular mechanics), and the mechanical effect of the Stone-Wales defects are scrutinized (via

ACS Paragon Plus Environment

2

Page 3 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

58

SMD). The atomistic simulations are performed using ReaxFF potential to describe interactions -

59

an empirical bond-order-dependent potential that allows for fully reactive atomistic scale

60

molecular dynamics simulations of chemical reactions, including bond rupture, in which

61

dissociation and reaction curves are obtained from fitting to quantum chemistry calculations

62

15

63

of many carbon systems. 16-37

64

First, to assess the stability of the nanothread and the ReaxFF potential, we compare the energy

65

of the system to the potential energy of a representative sheet of graphane (e.g., hydrogenated

66

graphene). Comparative analysis of the potential energies indicate a minimal difference in the

67

potential energy (less than 1%) per carbon atom, on the order of the energy change from rolling

68

graphene to a CNT and suggesting a stable allotrope (see Supporting Information). To confirm

69

our model is representative of previous synthetic efforts, we first assess the molecular structure

70

by calculating the radial distribution function (RDF), g(r), of a short equilibrium simulation (1

71

ns) at finite temperature (300K). We use a model thread with a length of approximately 8 nm.

72

The RDF is plotted in Fig. 1B, and matches previously published results.11 The observed nearest-

73

neighbor carbon–carbon distance of approx. 1.52 Å is characteristic of dominant sp3 bonding,

74

while the dominant peak at approx. 1.1 Å is consistent with the proposed 1:1 C/H stoichiometry.

75

Successive peaks are due to second nearest C-C and C-H correlations, as well as H-H pairs. The

76

peaks are more distinct in this model compared to experimental results11 due to the analysis of a

77

single thread – interthread pairwise correlations can result in slight variations and broadening

78

due to thread endcaps, cross-links, etc. The inclusion of a finite number of Stone-Wales defects

79

is necessary to produce the thread structure and representative RDF. To further explore the local

80

and global effects of the Stone-Wales defects, we consider four different structurally

14,

. Previous studies have successfully implemented ReaxFF for the behavior and characterization

ACS Paragon Plus Environment

3

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 4 of 25 Roman et al.

81

representative models containing 1 to 4 defects, resulting in only nominal variations in the

82

characterization. Thus, each of the models can be considered structurally representative.

83

However, energetic analysis indicates that adding a defect increases the potential energy of the

84

thread by approximately 12 kcal/mol.

85 86

We next apply uniaxial tensile strain along the thread axis to determine both stiffness and

87

maximum strength/strain. A thread with two Stone-Wales defects is chosen as the representative

88

model system (we note that two defects is necessary to produce a “unit segment” thread

89

structure; see Supporting Information). Using a quasi-static approach, cycles of incrementally

90

applied strain and minimization are undertaken (see Supporting Information). Subject to

91

strain, stress is calculated using a virial stress formulation, and can be plotted (see Fig. 2A). The

92

traditional definition of stress is difficult on quasi-one-dimensional materials due to the

93

ambiguity of cross-section definition. To alleviate this ambiguity, we also report an equivalent

94

1D stress (units of force), which are independent of cross-section, and thus easier to compare

95

across other 1D systems (see Supporting Information). The process is repeated until failure is

96

incurred, resulting in the maximum stress (134.3 GPa or 26.4 nN) and strain (14.9%). Once the

97

stress and strain are determined, the axial stiffness can then be simply calculated by:

98

= ∆⁄∆

99

To maintain the assumption of linear elasticity, the stiffness is determined using a linear fit of the

100

stress v. strain data up to  = 4%. This range was selected to fit the initial linear regime of stress-

101

strain, as there is a slight stiffening behavior beyond approximately 5% strain. We also note that

102

due to the ambiguity of the cross-section of the nanothread (e.g., a 0.5 nm diameter,

103

approximated by the van der Waal radii of opposite hydrogen atoms), the label of ‘‘modulus’’ is

(1)

ACS Paragon Plus Environment

4

Page 5 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

104

only intended as a placeholder for the tensile stiffness. The continuum interpretation of modulus,

105

E, is not reflective of such threads, and is merely used here as a convenient convention for

106

comparison. Using this approach, the stiffness of the nanothread is approximately 850 GPa (167

107

nN). In comparison to CNTs, we note that the stiffness of the nanothreads is ≈85% that of the

108

theoretical value of 1 TPa.38 As an alternate calculation of stiffness, we also plot the potential

109

energy density, U, versus strain (Fig. 2B). This enables a direct future comparison to ab initio

110

energy methods, such as DFT. The stiffness can then be expressed as:

111

=   ⁄ 

112

Fitting the MD data with all terms of a fourth-order polynomial (accounting for nonlinearities)

113

yields a tensile stiffness of E = 774 GPa (152 nN) at  = 2% (e.g., midpoint of the previous fit),

114

which is within 10% of the stiffness attained by the virial stress-strain approach, indicating

115

consistency between the two methods. Discrepancy can be attributed to the fact that only axial

116

stress was accounted for in the virial approach, neglecting higher order effects from shear

117

contributions, for example. A fourth-order fit was chosen by increasing the polynomial order

118

until sufficient fit (R2 > 0.98) was achieved.

(2)

119 120

As depicted, the ultimate stress of the thread is approximately 134.3 GPa or 26.4 nN. This is over

121

twice the strength of carbyne (single carbon chains) recently touted as one of the strongest new

122

materials39. We can also determine the specific strength, or tenacity, of the nanothread, which is

123

independent of the presumed cross-section and only dependent on the length of the thread, where

124

 =  ⁄, e.g., the ultimate stress normalized by density. This results in  = 4.13 × 10 N-

125

m/kg, potentially making these diamond nanothreads the strongest fabricated material by density

126

to date.

ACS Paragon Plus Environment

5

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 6 of 25 Roman et al.

127 128

While stiff in tension, we next assess the flexibility of the threads – do they behave more akin to

129

rods or cables? Through the equilibrium simulations, it is observed that the diamond nanothreads

130

are not an extremely rigid structure, and thus cannot be subjected to bending loads like a beam to

131

assess flexural response. Nor do the threads behave like flexible polymers, and thus

132

conformational methods (e.g., Kuhn length) cannot be applied to approximate bending rigidity.

133

As such, we determine out-of-plane bending stiffness/rigidity by introducing a virtual sticky

134

surface to deform along with the nanothread into curved configurations of various radii (Fig. 3A;

135

see Supporting Information). The bending stiffness is obtained by curve fitting with energy

136

versus curvature data points (Fig. 3B), using the following expression for elastic bending energy:

137

 =  "#  $

138

where U is the system strain energy, D is the bending stiffness, and κ is the prescribed curvature.

139

The resulting bending stiffness is calculated to be approximately 770 (kcal/mol)-Å, or 5.35 ×

140

10& N-m2. This makes the nanothread a somewhat rigid yet flexible molecule (a (5,5) CNT,

141

for example, has a rigidity on the order of 100,000 (kcal/mol)-Å, while carbyne has a rigidity of

142

30 to 80 (kcal/mol)-Å). In terms of persistence, where $' = "/)* , kB the Boltzmann constant,

143

T the temperature, and thus kBT the thermal energy, the diamond nanothread has a persistence

144

length of approximately 130 nm at 300K. Compared to polymers ( $' ≈ 1 nm) or double-

145

stranded DNA ($' ≈ 50 nm), diamond nanothreads are relatively rigid, while compared to CNTs

146

($' = 10 − 100 µm), they are relatively flexible. Thus, they behave more like a flexible (albeit

147

strong) cable rather than a rod or beam.

!

(3)

148

ACS Paragon Plus Environment

6

Page 7 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

149

One deviation between the structure of nanothreads and related nanotubes is the introduction of

150

Stone-Wales defects. These defects break the axial symmetry of the structure, and eliminate the

151

central hollow of the tube-like geometry. Observation of simulated tensile tests also indicated

152

that failure consistently occurred at the defect, suggesting a localization mechanism. To probe

153

the effect of such defects, we investigate the local strain under tensile loading, variation in local

154

stiffness, and predict overall thread behavior based on defect density.

155 156

First, the nanothread was divided into segments, to isolate the local strains associated with the

157

bonds adjacent the Stone-Wales defects. We plot the global stress versus local strain (Fig. 4A).

158

There is a clear difference between the “pristine” segments and the defective regions. The

159

pristine segments depict consistent stress-strain response, with an effective stiffness of

160

approximately 1089 GPa (214 nN). Note that these regions are effectively nanotubes, and it is no

161

surprise that this stiffness agrees with previous calculated values of such tubes.40 The strain at

162

failure is also on the order of 12%. In contrast, the defect regions are significantly more

163

compliant – fitting a modulus to the stress-strain data results in stiffness of 306 GPa (60 nN) and

164

347 GPa (68 nN) for the two defects, respectively. Note that that scatter in the plot is due to the

165

relatively small length and resolution of the strain measure (e.g., 1% strain is on the order of

166

hundredths of Ångstrom). Regardless, there is a clear trend – the strain is localized and much

167

greater at the defects.

168 169

To predict the mechanical response as a function of defect density, we consider the nanothread as

170

a system of serial springs (Fig. 4B). This has two significant implications: (1) the ultimate

171

strength is unaffected by number of defects and (2) the stiffness will decrease with number of

ACS Paragon Plus Environment

7

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 8 of 25 Roman et al.

172

defects. If the defect density is known, e.g., - = ./$/ , where n is the number of defects and L0

173

is the total length of the thread, theoretically we can predict the stiffness of the thread via a

174

simple rule of mixtures, where:

175

0.1 = 2

176

where Ld is the effective length of the defect region, and Ed and E0 the local stiffness of each

177

section. Assuming Ld ≈ 4.8 Å (measuring the approximate axial length across the defect area),

178

then 021 = 833 to 867 GPa (164 to 170 nN) depending on the value of - , which agrees well

179

with the molecular mechanics results for n = 2 (850 GPa or 167 nN).

345

65 47

+

047 345 1 ! 67 47

9

(4)

180 181

To confirm the model prediction, we implement steered molecular dynamics (SMD) to

182

investigate the stiffness and strength of the nanothread subject to applied loads at a low but finite

183

temperature (100K), rather than incremental affine strain (see Supporting Information). This

184

will facilitate any dynamic strain localization during applied loading and ultimate failure event,

185

and more accurately represent experimental strength measures. Rate dependence was checked by

186

varying loading velocity (see Supporting Information). Threads with 2, 3, and 4 defects were ruptured

187

via SMD, and the stress-strain response extracted from the force-displacement results (see Fig.

188

4C).

189 190

We find that the threads all break at the same load, approximately 128 GPa (25.1 nN). This is

191

within 5% of the measured molecular mechanics strength of 134.3 GPa (26.4 nN) indicating a

192

slight weakening due to temperature. From a statistical perspective, one may expect a weakening

193

with the introduction of more defects, as failure may occur at a greater number of locations,

194

assuming a distribution of stresses and local variance. This would also result in a length

ACS Paragon Plus Environment

8

Page 9 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

195

dependence on strength. Due to the limited number of defects (n≤4) and constant length, this

196

behavior was not observed. However, this could also be due to the limited temperature of the

197

SMD simulations, which limits stress variation across the thread. Such statistical measures are

198

beyond the scope of the current first-order characterization. Moreover, while the SMD load is

199

applied in the axial direction, the thread is able to move out of alignment, which may also cause

200

additional weakening. In consideration of these effects, the 5% deviation is deemed acceptable.

201

In terms of stiffness, we see a consistent decrease with an increase in defect density, with

202

modulii of 808 GPa (158 nN) 763 GPa (150 nN) and 665 GPa (131 nN) for n = 2, 3, and 4

203

respectively (measured between strains of 3%-8%). Note again, there is a slight increase in

204

compliance attributed to temperature effects (808 GPa compared to 850 GPa; a 4.9% difference).

205

From the theoretical prediction, 031 = 746 to 786 GPa while 041 = 675 to 720 GPa,

206

consistent with the observed reduction in magnitudes (see Fig 5). With the introduction of the

207

different numbers of defects, the stiffness present a degradation as a function of the defect

208

density. We also see an increase in ultimate strain (strain at failure), between 19.6% and 22.7%

209

(note that the increase in ultimate strain is due to initial straightening of the thread due to the

210

curvature imposed by temperature effects; the initial gauge length was measured relative to the

211

slightly curved equilibrated structure, not a straight thread).

212 213

Of interest, we also track the bond-order (BO) evolution of the carbon-carbon bonds within the

214

nanothread during extension (see Supporting Information). From BO inspection, we can make

215

two conclusions: First, the system maintains its sp3 character throughout the tensile loading until

216

failure, which could be key in exploiting electrical or optical properties of the nanothreads.

217

Second, only the bonds primarily aligned with the axis of extension (e.g., along the thread

ACS Paragon Plus Environment

9

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 10 of 25 Roman et al.

218

length) undergo a reduction in BO, implying tension, whereas the bonds transverse to the axis

219

undergo a nominal increase in BO, implying slight compression. Thus, the thread acts marginally

220

like a nanotruss, wherein the transverse bonds serve as stiffening members rather than a majority

221

of the bonds under tension. This can further explicate (a) the localization of failure on the defect

222

region (which is unreinforced by transverse bonds), and (b) the deviation between virial

223

approach and energy approach in stiffness.

224 225

In summary, we investigated the mechanical properties of diamond-like nanothreads of carbon,

226

to assess the potential high strength and stiffness, which allows identifying the imminent benefits

227

and mechanical limits of its application in different areas. It was predicted that, due to the

228

diamond-like structure, these nanothreads may have exhibited extraordinary properties such as

229

strength and stiffness higher than that carbon nanotubes or conventional high-strength polymers.

230

We have shown an axial stiffness on the order of nanotubes (approximately 665 to 850 GPa or

231

131 to 167 nN) depending on defect density and characterization method. A specific strength

232

( 3.94 × 10 to 4.13 × 10 N-m/kg) exceeding all known fabricated materials. This specific

233

property of the diamond nanothread is particularly encouraging, making it a material that can be

234

featured as the solution to the ever-challenging space elevator problem. Moreover, the strength is

235

independent of the number of Stone-Wales defects, perhaps facilitating the synthesis of large-

236

scale thread systems (e.g., bundles or fabrics) without significant loss of capacity (unlike carbon

237

nanotubes). In 2D and 3D crystalline materials, it is well-known that the density of defects

238

affects the plasticity and brittle-ductile transitions. Here, the nanothread is different because (a) it

239

is inherently 1D and (b) the defects are not an error in an otherwise ordered crystalline structure,

240

but rather necessary to define the nanothread (otherwise, it would be a nanotube), and the

ACS Paragon Plus Environment

10

Page 11 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

241

molecular structure is essentially a Stone-Wales defect in geometry alone. Finally, we show the

242

thread can be represented by a simple springs-in-series model and, while the strength remains the

243

same, the stiffness can be predicted by the linear defect density (ρd=n/L0). Due to the linear

244

spring-like arrangement, one defect must carry the load just as 2, 3, or 4 defects would, thus the

245

“brittleness” in terms of sudden failure does not change. Indeed, the strength drop due to defects

246

here is not the same as a strength drop due to stress concentrations, as usually encountered with a

247

defected system – due to the linear arrangement, all sections (with defect or defect-free) must

248

carry the same load. Thus, additional defects do not weaken the system, but define the upper

249

bound of strength. In terms of plasticity, we have also shown that the ultimate strain increases

250

when the number of defects increases.

251 252

These nanothreads may be the first member of a new class of diamond-like nanomaterials based

253

on a strong tetrahedral core, and determining the mechanical limits of such systems is necessary

254

to aid application development. While strong in tension, the relatively low bending rigidity may

255

facilitate the design of thread bundles or weaved systems, lacking the mechanical persistence of

256

CNTs. The lack of crystallinity may enable allowances of defects with suffering mechanical

257

degradation. Functionalized nanothreads offer the potential for improved load transfer through

258

covalent bonding, efficiently transferring their mechanical strength to a surrounding matrix and

259

thus allowing for technological exploitation in fibres or fabrics. Once again, carbon has provided

260

yet another exemplary allotrope to explore and engineer.

261 262 263 264 265 266

Acknowledgements R.R., K.K., and S.W.C. acknowledge generous support from NEU’s CEE Department. The calculations and the analysis were carried out using a parallel LINUX cluster at NEU’s Laboratory for Nanotechnology In Civil Engineering (NICE). Visualization has been carried out using the VMD visualization package.41

ACS Paragon Plus Environment

11

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 12 of 25 Roman et al.

267

Citations

268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316

1. Baughman, R. H.; Eckhardt, H.; Kertesz, M. The Journal of Chemical Physics 1987, 87, (11), 6687. 2. Baughman, R. H.; Zakhidov, A. A.; de Heer, W. A. Science 2002, 297, (5582), 787-792. 3. Cranford, S.; Buehler, M. J. Model Simul Mater Sc 2011, 19, (5). 4. Cranford, S. W.; Brommer, D. B.; Buehler, M. J. Nanoscale 2012, 4, (24), 7797-7809. 5. Cranford, S. W.; Buehler, M. J. Carbon 2011, 49, (13), 4111-4121. 6. Geim, A. K.; Novoselov, K. S. Nat Mater 2007, 6, (3), 183-191. 7. Hirsch, A. Nat Mater 2010, 9, (11), 868-871. 8. Jones, R. O.; Seifert, G. Physical Review Letters 1997, 79, (3), 443-446. 9. Kocsis, A. J.; Yedama, N. A. R.; Cranford, S. W. Nanotechnology 2014, 25, (33), 335709. 10. Mirzaeifar, R.; Qin, Z.; Buehler, M. J. Nanotechnology 2014, 25, (37), 371001. 11. Fitzgibbons, T. C.; Guthrie, M.; Xu, E.-s.; Crespi, V. H.; Davidowski, S. K.; Cody, G. D.; Alem, N.; Badding, J. V. Nat Mater 2014, advance online publication. 12. Stojkovic, D.; Zhang, P. H.; Crespi, V. H. Physical Review Letters 2001, 87, (12). 13. Yang, Y. L.; Xu, X. M. Comp Mater Sci 2012, 61, 83-88. 14. van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; Goddard III, W. A. The Journal of Physical Chemistry A 2001, 105, (41), 9396–9409. 15. Chenoweth, K.; van Duin, A. C. T.; Goddard, W. A. The Journal of Physical Chemistry A 2008, 112, (5), 1040-1053. 16. Srinivasan, S. G.; van Duin, A. C. T.; Ganesh, P. The Journal of Physical Chemistry A 2015, 119, (4), 571-580. 17. Page, A. J.; Moghtaderi, B. The Journal of Physical Chemistry A 2009, 113, (8), 1539-1547. 18. Zaminpayma, E.; Nayebi, P. Physica B: Condensed Matter 2015, 459, (0), 29-35. 19. Goverapet Srinivasan, S.; van Duin, A. C. T. Carbon 2015, 82, (0), 314-326. 20. Zhang, C.; Wen, Y.; Xue, X. ACS Applied Materials & Interfaces 2014, 6, (15), 12235-12244. 21. Paupitz, R.; Junkermeier, C. E.; van Duin, A. C. T.; Branicio, P. S. Physical Chemistry Chemical Physics 2014, 16, (46), 25515-25522. 22. Kwan, K.; Cranford, S. W. Applied Physics Letters 2014, 104, (20), 203101. 23. Bhoi, S.; Banerjee, T.; Mohanty, K. Fuel 2014, 136, (0), 326-333. 24. Berdiyorov, G. R.; Milošević, M. V.; Peeters, F. M.; van Duin, A. C. T. Physica B: Condensed Matter 2014, 455, (0), 60-65. 25. Poovathingal, S.; Schwartzentruber, T. E.; Srinivasan, S. G.; van Duin, A. C. T. The Journal of Physical Chemistry A 2013, 117, (13), 2692-2703. 26. Neyts, E. C.; Ostrikov, K.; Han, Z. J.; Kumar, S.; van Duin, A. C. T.; Bogaerts, A. Physical Review Letters 2013, 110, (6), 065501. 27. Liu, X. Y.; Wang, F. C.; Park, H. S.; Wu, H. A. Journal of Applied Physics 2013, 114, (5), 054313. 28. Huang, L.; Seredych, M.; Bandosz, T. J.; van Duin, A. C. T.; Lu, X.; Gubbins, K. E. The Journal of Chemical Physics 2013, 139, (19), 194707. 29. Santos, R. P. B. d.; Perim, E.; Autreto, P. A. S.; Gustavo, B.; Galvão, D. S. Nanotechnology 2012, 23, (46), 465702. 30. Kim, K.; Artyukhov, V. I.; Regan, W.; Liu, Y.; Crommie, M. F.; Yakobson, B. I.; Zettl, A. Nano Letters 2011, 12, (1), 293-297. 31. Compton, O. C.; Cranford, S. W.; Putz, K. W.; An, Z.; Brinson, L. C.; Buehler, M. J.; Nguyen, S. T. ACS Nano 2011, 6, (3), 2008-2019. 32. Sen, D.; Novoselov, K. S.; Reis, P. M.; Buehler, M. J. Small 2010, 6, (10), 1108-1116. 33. Neyts, E. C.; Shibuta, Y.; van Duin, A. C. T.; Bogaerts, A. ACS Nano 2010, 4, (11), 6665-6672. 34. Bagri, A.; Grantab, R.; Medhekar, N. V.; Shenoy, V. B. The Journal of Physical Chemistry C 2010, 114, (28), 12053-12061.

ACS Paragon Plus Environment

12

Page 13 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

317 318 319 320 321 322 323 324 325 326 327 328 329

Roman et al.

35. Zhang, L.; Zybin, S. V.; van Duin, A. C. T.; Dasgupta, S.; Goddard, W. A.; Kober, E. M. The Journal of Physical Chemistry A 2009, 113, (40), 10619-10640. 36. Flores, M. Z. S.; Autreto, P. A. S.; Legoas, S. B.; Galvao, D. S. Nanotechnology 2009, 20, (46), 465704. 37. Ganesh, P.; Kent, P. R. C.; Mochalin, V. Journal of Applied Physics 2011, 110, (7), 073506. 38. Bao, W. X.; Zhu, C. C.; Cui, W. Z. Physica B 2004, 352, (1-4), 156-163. 39. Liu, M. J.; Artyukhov, V. I.; Lee, H.; Xu, F. B.; Yakobson, B. I. ACS Nano 2013, 7, (11), 1007510082. 40. Krishnan, A.; Dujardin, E.; Ebbesen, T. W.; Yianilos, P. N.; Treacy, M. M. J. Phys Rev B 1998, 58, (20), 14013-14019. 41. Humphrey, W.; Dalke, A.; Schulten, K. J Mol Graphics 1996, 14, (1), 33. Figures and Figure Captions

330 331 332 333 334 335 336 337 338 339

Figure 1: Diamond nanothread structure. A. Simulation snapshot of diamond nanothread model; approximately 8 nm in length with two Stone-Wales defects (see inset) located approximately 2 nm from each edge. From a modeling perspective it can be considered as a variation of a hydrogenated (3,0) nanotube with Stone-Wales defects. B. Radial distribution function (RDF), g(r), of a short equilibrium simulation (1 ns) of diamond nanothread at finite temperature (300K). The RDF and matches previously published results11 with key peaks representing the 1:1 carbon-hydrogen stoichiometry and thread structure. Multiple lines with deviation represent multiple runs with variation in number of defects (n = 1,2,3 and 4), indicating negligible structural change with defect density.

ACS Paragon Plus Environment

13

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Page 14 of 25 Roman et al.

340 341

ACS Paragon Plus Environment

14

Page 15 of 25

Nano Letters Mech. Prop. of Dia. Nanothreads

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

Roman et al.

342 343 344 345 346 347 348 349

Figure 2: Molecular mechanics tensile test. A. Virial stress versus strain for quasi-static incremental strain loading (strain increments of 0.0001 followed by energy minimization). Fitted stiffness (ε