ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/NPthiols/NPthiols.tex
Revision: 4384
Committed: Wed Oct 28 14:54:40 2015 UTC (8 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 48355 byte(s)
Log Message:
Latest edits

File Contents

# Content
1 %% ****** Start of file template.aps ****** %
2 %%
3 %%
4 %% This file is part of the APS files in the REVTeX 4 distribution.
5 %% Version 4.0 of REVTeX, August 2001
6 %%
7 %%
8 %% Copyright (c) 2001 The American Physical Society.
9 %%
10 %% See the REVTeX 4 README file for restrictions and more information.
11 %%
12 %
13 % This is a template for producing manuscripts for use with REVTEX 4.0
14 % Copy this file to another name and then work on that file.
15 % That way, you always have this original template file to use.
16 %
17 % Group addresses by affiliation; use superscriptaddress for long
18 % author lists, or if there are many overlapping affiliations.
19 % For Phys. Rev. appearance, change preprint to twocolumn.
20 % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21 % Add 'draft' option to mark overfull boxes with black boxes
22 % Add 'showpacs' option to make PACS codes appear
23 %\documentclass[aps,jcp,twocolumn,showpacs,superscriptaddress,groupedaddress]{revtex4} % for review and submission
24 \documentclass[aps,jcp,preprint,showpacs,superscriptaddress,groupedaddress]{revtex4-1} % for double-spaced preprint
25 \usepackage{graphicx} % needed for figures
26 \usepackage{bm} % for math
27 \usepackage{amssymb} % for math
28 \usepackage{times}
29 \usepackage[version=3]{mhchem}
30 \usepackage{lineno}
31
32 \begin{document}
33
34 \title{Interfacial Thermal Conductance of Thiolate-Protected
35 Gold Nanospheres}
36 \author{Kelsey M. Stocker}
37 \author{Suzanne M. Neidhart}
38 \author{J. Daniel Gezelter}
39 \email{gezelter@nd.edu}
40 \affiliation{Department of Chemistry and Biochemistry, University of
41 Notre Dame, Notre Dame, IN 46556}
42
43 \begin{abstract}
44 Molecular dynamics simulations of thiolate-protected and solvated
45 gold nanoparticles were carried out in the presence of a
46 non-equilibrium heat flux between the solvent and the core of the
47 particle. The interfacial thermal conductance ($G$) was computed for
48 these interfaces, and the behavior of the thermal conductance was
49 studied as a function of particle size, ligand flexibility, and
50 ligand chain length. In all cases, thermal conductance of the
51 ligand-protected particles was higher than the bare metal--solvent
52 interface. A number of mechanisms for the enhanced conductance were
53 investigated, including thiolate-driven corrugation of the metal
54 surface, solvent ordering at the interface, solvent-ligand
55 interpenetration, and ligand ordering relative to the particle
56 surface. Only the smallest particles exhibited significant
57 corrugation. All ligands permitted substantial solvent-ligand
58 interpenetration, and ligand chain length has a significant
59 influence on the orientational ordering of interfacial solvent.
60 Solvent -- ligand vibrational overlap, particularly in the low
61 frequency range ($< 80 \mathrm{cm}^{-1}$) was significantly altered
62 by ligand rigidity, and had direct influence on the interfacial
63 thermal conductance.
64 \end{abstract}
65
66 \pacs{}
67 \keywords{}
68 \maketitle
69
70 \section{Introduction}
71
72 Heat transport across various nanostructured interfaces has been the
73 subject of intense experimental
74 interest,\cite{Wilson:2002uq,Ge:2004yg,Shenogina:2009ix,Wang10082007,Schmidt:2008ad,Juve:2009pt,Alper:2010pd,Harikrishna:2013ys}
75 and the interfacial thermal conductance, $G$, is the principal
76 quantity of interest for understanding interfacial heat
77 transport.\cite{Cahill:2003fk} Because nanoparticles have a
78 significant fraction of their atoms at the particle / solvent
79 interface, the chemical details of these interfaces govern the thermal
80 transport properties. For ligand-protected nanoparticles, there may
81 be three distinct heat transfer processes: (1) from the particles to
82 the ligands, (2) vibrational energy tranfer along the length of the
83 ligand, followed by (3) heat transport from the ligand to the
84 surrounding solvent.\cite{Ge:2006kx}
85
86 Heat transport at the gold-alkylthiolate-solvent interface has been
87 previously explored both through molecular dynamics simulations and
88 via time domain
89 thermoreflectance.\cite{Kikugawa:2009vn,Kuang:2011ef,Stocker:2013cl,Tian:2015uq}
90 Most of these studies have found that alkylthiolates enhance the
91 thermal conductance to the solvent, and that the vibrational overlap
92 provided by the chemically-bound ligand species plays a role in this
93 enhancement.
94
95 Reverse nonequilibrium molecular dynamics (RNEMD) methods have been
96 previously applied to calculate the thermal conductance at flat (111)
97 metal / organic solvent interfaces that had been chemically protected
98 by varying coverages of alkanethiolate groups.\cite{Kuang:2011ef}
99 These simulations suggested an explanation for the increased thermal
100 conductivity at alkanethiol-capped metal surfaces compared with bare
101 metal interfaces. Specifically, the chemical bond between the metal
102 and the ligand introduces a vibrational overlap that is not present
103 without the protecting group, and the overlap between the vibrational
104 spectra (metal to ligand, ligand to solvent) provides a mechanism for
105 rapid thermal transport across the interface. The simulations also
106 suggested that this phenomenon is a non-monotonic function of the
107 fractional coverage of the surface, as moderate coverages allow energy
108 transfer to solvent molecules that come into close contact with the
109 ligands.
110
111 Similarly, simulations of {\it mixed-chain} alkylthiolate surfaces
112 showed that solvent trapped close to the interface can be efficient at
113 moving thermal energy away from the surface.\cite{Stocker:2013cl}
114 Trapped solvent molecules that were orientationally aligned with
115 nearby ligands were able to increase the thermal conductance of the
116 interface. This indicates that the ligand-to-solvent vibrational
117 energy transfer is a key feature for increasing particle-to-solvent
118 thermal conductance.
119
120 Recently, we extended RNEMD methods for use in non-periodic geometries
121 by creating scaling/shearing moves between concentric regions of a
122 simulation.\cite{Stocker:2014qq} In this work, we apply this
123 non-periodic variant of RNEMD to investigate the role that {\it
124 curved} nanoparticle surfaces play in heat and mass transport. On
125 planar surfaces, we discovered that orientational ordering of surface
126 protecting ligands had a large effect on the heat conduction from the
127 metal to the solvent. Smaller nanoparticles have high surface
128 curvature that creates gaps in well-ordered self-assembled monolayers,
129 and the effect of those gaps on the thermal conductance is unknown.
130
131 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
132 % INTERFACIAL THERMAL CONDUCTANCE OF METALLIC NANOPARTICLES
133 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134 %\section{Interfacial Thermal Conductance of Metallic Nanoparticles}
135
136 For a solvated nanoparticle, it is possible to define a critical value
137 for the interfacial thermal conductance,
138 \begin{equation}
139 G_c = \frac{3 C_s \Lambda_s}{R C_p}
140 \end{equation}
141 which depends on the solvent heat capacity, $C_s$, solvent thermal
142 conductivity, $\Lambda_s$, particle radius, $R$, and nanoparticle heat
143 capacity, $C_p$.\cite{Wilson:2002uq} In the limit of infinite
144 interfacial thermal conductance, $G \gg G_c$, cooling of the
145 nanoparticle is limited by the solvent properties, $C_s$ and
146 $\Lambda_s$. In the opposite limit, $G \ll G_c$, the heat dissipation
147 is controlled by the thermal conductance of the particle / fluid
148 interface. It is this regime with which we are concerned, where
149 properties of ligands and the particle surface may be tuned to
150 manipulate the rate of cooling for solvated nanoparticles. Based on
151 estimates of $G$ from previous simulations as well as experimental
152 results for solvated nanostructures, gold nanoparticles solvated in
153 hexane are in the $G \ll G_c$ regime for radii smaller than 40 nm. The
154 particles included in this study are more than an order of magnitude
155 smaller than this critical radius, so the heat dissipation should be
156 controlled entirely by the surface features of the particle / ligand /
157 solvent interface.
158
159 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160 % STRUCTURE OF SELF-ASSEMBLED MONOLAYERS ON NANOPARTICLES
161 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
162 \subsection{Structures of Self-Assembled Monolayers on Nanoparticles}
163
164 Though the ligand packing on planar surfaces has been characterized
165 for many different ligands and surface facets, it is not obvious
166 \emph{a priori} how the same ligands will behave on the highly curved
167 surfaces of spherical nanoparticles. Thus, as new applications of
168 ligand-stabilized nanostructures have been proposed, the structure and
169 dynamics of ligands on metallic nanoparticles have been studied using
170 molecular simulation,\cite{Henz:2008qf} NMR, XPS, FTIR,
171 calorimetry, and surface
172 microscopies.\cite{Badia1996:2,Badia1996,Badia1997:2,Badia1997,Badia2000}
173 Badia, \textit{et al.} used transmission electron microscopy to
174 determine that alkanethiol ligands on gold nanoparticles pack
175 approximately 30\% more densely than on planar Au(111)
176 surfaces.\cite{Badia1996:2} Subsequent experiments demonstrated that
177 even at full coverages, surface curvature creates voids between linear
178 ligand chains that can be filled via interdigitation of ligands on
179 neighboring particles.\cite{Badia1996} The molecular dynamics
180 simulations of Henz, \textit{et al.} indicate that at low coverages,
181 the thiolate alkane chains will lie flat on the nanoparticle
182 surface\cite{Henz:2008qf} Above 90\% coverage, the ligands
183 stand upright and recover the rigidity and tilt angle displayed on
184 planar facets. Their simulations also indicate a high degree of mixing
185 between the thiolate sulfur atoms and surface gold atoms at high
186 coverages.
187
188 In this work, thiolated gold nanospheres were modeled using a united
189 atom force field and non-equilibrium molecular dynamics. Gold
190 nanoparticles with radii ranging from 10 - 25 \AA\ were created from a
191 bulk fcc lattice. These particles were passivated with a 50\%
192 coverage -- based on (compared with the coverage densities reported by
193 Badia \textit{et al.}) of a selection of thiolates. Three
194 straight-chain thiolates of varying chain lengths and rigidities were
195 utilized. These are summarized in Fig. \ref{fig:structures}. The
196 passivated particles were then solvated in hexane. Details on the
197 united atom force field are given below and in the supporting
198 information.
199
200 \begin{figure}
201 \includegraphics[width=\linewidth]{figures/structures}
202 \caption{Topologies of the thiolate capping agents and solvent
203 utilized in the simulations. The chemically-distinct sites (S, \ce{CH}
204 \ce{CH2}, \ce{CH3}, and \ce{CHa}) are treated as united atoms. Most
205 parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
206 \protect\cite{TraPPE-UA.alkanes} and
207 \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
208 atoms were adapted from references
209 \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
210 \protect\cite{hautman:4994}.}
211 \label{fig:structures}
212 \bibpunct{[}{]}{,}{n}{}{,}
213 \end{figure}
214
215
216 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
217 % COMPUTATIONAL DETAILS
218 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
219 \section{Computational Details}
220
221 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
222 % NON-PERIODIC VSS-RNEMD METHODOLOGY
223 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
224 \subsection{Creating a thermal flux between particles and solvent}
225
226 The non-periodic variant of VSS-RNEMD\cite{Stocker:2014qq} applies a
227 series of velocity scaling and shearing moves at regular intervals to
228 impose a flux between two concentric spherical regions. To impose a
229 thermal flux between the shells (without an accompanying angular
230 shear), we solve for scaling coefficients $a$ and $b$,
231 \begin{eqnarray}
232 a = \sqrt{1 - \frac{q_r \Delta t}{K_a - K_a^\mathrm{rot}}}\\ \nonumber\\
233 b = \sqrt{1 + \frac{q_r \Delta t}{K_b - K_b^\mathrm{rot}}}
234 \end{eqnarray}
235 at each time interval. These scaling coefficients conserve total
236 kinetic energy and angular momentum subject to an imposed heat rate,
237 $q_r$. The coefficients also depend on the instantaneous kinetic
238 energy, $K_{\{a,b\}}$, and the total rotational kinetic energy of each
239 shell, $K_{\{a,b\}}^\mathrm{rot} = \sum_i m_i \left( \mathbf{v}_i
240 \times \mathbf{r}_i \right)^2 / 2$.
241
242 The scaling coefficients are determined and the velocity changes are
243 applied at regular intervals,
244 \begin{eqnarray}
245 \mathbf{v}_i \leftarrow a \left ( \mathbf{v}_i - \left < \omega_a \right > \times \mathbf{r}_i \right ) + \left < \omega_a \right > \times \mathbf{r}_i~~\:\\
246 \mathbf{v}_j \leftarrow b \left ( \mathbf{v}_j - \left < \omega_b \right > \times \mathbf{r}_j \right ) + \left < \omega_b \right > \times \mathbf{r}_j.
247 \end{eqnarray}
248 Here $\left < \omega_a \right > \times \mathbf{r}_i$ is the
249 contribution to the velocity of particle $i$ due to the overall
250 angular velocity of the $a$ shell. In the absence of an angular
251 momentum flux, the angular velocity $\left < \omega_a \right >$ of the
252 shell is nearly 0 and the resultant particle velocity is a nearly
253 linear scaling of the initial velocity by the coefficient $a$ or $b$.
254
255 Repeated application of this thermal energy exchange yields a radial
256 temperature profile for the solvated nanoparticles that depends
257 linearly on the applied heat rate, $q_r$. Similar to the behavior in
258 the slab geometries, the temperature profiles have discontinuities at
259 the interfaces between dissimilar materials. The size of the
260 discontinuity depends on the interfacial thermal conductance, which is
261 the primary quantity of interest.
262
263 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264 % CALCULATING TRANSPORT PROPERTIES
265 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
266 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267 % INTERFACIAL THERMAL CONDUCTANCE
268 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
269 \subsection{Interfacial Thermal Conductance}
270
271 As described in earlier work,\cite{Stocker:2014qq} the thermal
272 conductance of each spherical shell may be defined as the inverse
273 Kapitza resistance of the shell. To describe the thermal conductance
274 of an interface of considerable thickness -- such as the ligand layers
275 shown here -- we can sum the individual thermal resistances of each
276 concentric spherical shell to arrive at the inverse of the total
277 interfacial thermal conductance. In slab geometries, the intermediate
278 temperatures cancel, but for concentric spherical shells, the
279 intermediate temperatures and surface areas remain in the final sum,
280 requiring the use of a series of individual resistance terms:
281
282 \begin{equation}
283 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
284 T_i\right) 4 \pi r_i^2.
285 \end{equation}
286
287 The longest ligand considered here is in excess of 15 \AA\ in length,
288 and we use 10 concentric spherical shells to describe the total
289 interfacial thermal conductance of the ligand layer.
290
291 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
292 % FORCE FIELDS
293 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
294 \subsection{Force Fields}
295
296 Throughout this work, gold -- gold interactions are described by the
297 quantum Sutton-Chen (QSC) model.\cite{Qi:1999ph} Previous
298 work\cite{Kuang:2011ef} has demonstrated that the electronic
299 contributions to heat conduction (which are missing from the QSC
300 model) across heterogeneous metal / non-metal interfaces are
301 negligible compared to phonon excitation, which is captured by the
302 classical model. The hexane solvent is described by the TraPPE united
303 atom model,\cite{TraPPE-UA.alkanes} where sites are located at the
304 carbon centers for alkyl groups. The TraPPE-UA model for hexane
305 provides both computational efficiency and reasonable accuracy for
306 bulk thermal conductivity values. Bonding interactions were used for
307 intra-molecular sites closer than 3 bonds. Effective Lennard-Jones
308 potentials were used for non-bonded interactions.
309
310 The TraPPE-UA force field includes parameters for thiol
311 molecules\cite{TraPPE-UA.thiols} as well as unsaturated and aromatic
312 carbon sites.\cite{TraPPE-UA.alkylbenzenes} These were used for the
313 thiolate molecules in our simulations, and missing parameters for the
314 ligands were supplemented using fits described in the supporting
315 information. Bonds are typically rigid in TraPPE-UA, so although
316 equilibrium bond distances were taken from TraPPE-UA, flexible bonds
317 were allowed bond stretching spring constants from the OPLS-AA force
318 field.\cite{Jorgensen:1996sf}
319
320 To derive suitable parameters for the thiolates adsorbed on Au(111)
321 surfaces, we adopted the S parameters from Luedtke and
322 Landman\cite{landman:1998} and modified the parameters for the CTS
323 atom to maintain charge neutrality in the molecule.
324
325 Other interactions between metal (Au) and non-metal atoms were adapted
326 from an adsorption study of alkyl thiols on gold surfaces by Vlugt,
327 \textit{et al.}\cite{vlugt:cpc2007154} They fit an effective pair-wise
328 Lennard-Jones form of potential parameters for the interaction between
329 Au and pseudo-atoms CH$_x$ and S based on a well-established and
330 widely-used effective potential of Hautman and Klein for the Au(111)
331 surface.\cite{hautman:4994}
332
333 All additional terms to represent thiolated alkenes and conjugated
334 ligand moieties were parameterized as part of this work and are
335 available in the supporting information.
336
337 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
338 % SIMULATION PROTOCOL
339 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
340 \subsection{Simulation Protocol}
341
342 Gold nanospheres with radii ranging from 10 - 25 \AA\ were created
343 from a bulk fcc lattice and were thermally equilibrated prior to the
344 addition of ligands. A 50\% coverage of ligands (based on coverages
345 reported by Badia, \textit{et al.}\cite{Badia1996:2}) were placed on
346 the surface of the equilibrated nanoparticles using
347 Packmol\cite{packmol}. We have chosen three lengths for the
348 straight-chain ligands, $C_4$, $C_8$, and $C_{12}$, differentiated by
349 the number of carbons in the chains. Additionally, to explore the
350 effects of ligand flexibility, we have used three levels of ligand
351 ``stiffness''. The most flexible chain is a fully saturated
352 alkanethiolate, while moderate rigidity is introduced using an alkene
353 thiolate with one double bond in the penultimate (solvent-facing)
354 carbon-carbon location. The most rigid ligands are fully-conjugated
355 chains where all of the carbons are represented with conjugated (aryl)
356 united-atom carbon atoms (CHar or terminal CH2ar).
357
358 The nanoparticle / ligand complexes were thermally equilibrated to
359 allow for ligand conformational flexibility. Packmol was then used to
360 solvate the structures inside a spherical droplet of hexane. The
361 thickness of the solvent layer was chosen to be at least 1.5$\times$
362 the combined radius of the nanoparticle / ligand structure. The fully
363 solvated system was equilibrated for at least 1 ns using the Langevin
364 Hull to apply 50 atm of pressure and a target temperature of 250
365 K.\cite{Vardeman2011} Typical system sizes ranged from 18,310 united
366 atom sites for the 10 \AA\ particles with $C_4$ ligands to 89,490
367 sites for the 25 \AA\ particles with $C_{12}$ ligands. Figure
368 \ref{fig:NP25_C12h1} shows one of the solvated 25 \AA\ nanoparticles
369 passivated with the $C_{12}$ alkane thiolate ligands.
370
371 \begin{figure}
372 \includegraphics[width=\linewidth]{figures/NP25_C12h1}
373 \caption{A 25 \AA\ radius gold nanoparticle protected with a
374 half-monolayer of TraPPE-UA dodecanethiolate (C$_{12}$) ligands
375 and solvated in TraPPE-UA hexane. The interfacial thermal
376 conductance is computed by applying a kinetic energy flux between
377 the nanoparticle and an outer shell of solvent.}
378 \label{fig:NP25_C12h1}
379 \end{figure}
380
381 Once equilibrated, thermal fluxes were applied for 1 ns, until stable
382 temperature gradients had developed. Systems were run under moderate
383 pressure (50 atm) with an average temperature (250K) that maintained a
384 compact solvent cluster and avoided formation of a vapor layer near
385 the heated metal surface. Pressure was applied to the system via the
386 non-periodic Langevin Hull.\cite{Vardeman2011} However, thermal
387 coupling to the external temperature bath was removed to avoid
388 interference with the imposed RNEMD flux.
389
390 \begin{figure}
391 \includegraphics[width=\linewidth]{figures/temp_profile}
392 \caption{Radial temperature profile for a 25 \AA\ radius
393 particle protected with a 50\% coverage of TraPPE-UA
394 butanethiolate (C$_4$) ligands and solvated in TraPPE-UA
395 hexane. A kinetic energy flux is applied between RNEMD
396 region A and RNEMD region B. The size of the temperature
397 discontinuity at the interface is governed by the
398 interfacial thermal conductance.}
399 \label{fig:temp_profile}
400 \end{figure}
401
402 Because the method conserves \emph{total} angular momentum and energy,
403 systems which contain a metal nanoparticle embedded in a significant
404 volume of solvent will still experience nanoparticle diffusion inside
405 the solvent droplet. To aid in measuring an accurate temperature
406 profile for these systems, a single gold atom at the origin of the
407 coordinate system was assigned a mass $10,000 \times$ its original
408 mass. The bonded and nonbonded interactions for this atom remain
409 unchanged and the heavy atom is excluded from the RNEMD velocity
410 scaling. The only effect of this gold atom is to effectively pin the
411 nanoparticle at the origin of the coordinate system, thereby
412 preventing translational diffusion of the nanoparticle due to Brownian
413 motion.
414
415 To provide statistical independence, five separate configurations were
416 simulated for each particle radius and ligand. The structures were
417 unique, starting at the point of ligand placement, in order to sample
418 multiple surface-ligand configurations.
419
420
421 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
422 % EFFECT OF PARTICLE SIZE
423 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
424 \section{Results}
425
426 We modeled four sizes of nanoparticles ($R =$ 10, 15, 20, and 25
427 \AA). The smallest particle size produces the lowest interfacial
428 thermal conductance values for most of the of protecting groups
429 (Fig. \ref{fig:NPthiols_G}). Between the other three sizes of
430 nanoparticles, there is no systematic dependence of the interfacial
431 thermal conductance on the nanoparticle size. It is likely that the
432 differences in local curvature of the nanoparticle sizes studied here
433 do not disrupt the ligand packing and behavior in drastically
434 different ways.
435
436 \begin{figure}
437 \includegraphics[width=\linewidth]{figures/G3}
438 \caption{Interfacial thermal conductance ($G$) values for 4
439 sizes of solvated nanoparticles that are bare or protected with
440 a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$ thiolate
441 ligands. Ligands of different flexibility are shown in separate
442 panels. The middle panel indicates ligands which have a single
443 carbon-carbon double bond in the penultimate position.}
444 \label{fig:NPthiols_G}
445 \end{figure}
446
447 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
448 % EFFECT OF LIGAND CHAIN LENGTH
449 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450
451 Unlike our previous study of varying thiolate ligand chain lengths on
452 planar Au(111) surfaces, the interfacial thermal conductance of
453 ligand-protected nanospheres exhibits a distinct dependence on the
454 ligand identity. A half-monolayer coverage of ligands yields
455 interfacial conductance that is strongly dependent on both ligand
456 length and flexibility.
457
458 There are many factors that could be playing a role in the
459 ligand-dependent conductuance. The sulfur-gold interaction is
460 particularly strong, and the presence of the ligands can easily
461 disrupt the crystalline structure of the gold at the surface of the
462 particles, providing more efficient scattering of phonons into the
463 ligand / solvent layer. This effect would be particularly important at
464 small particle sizes.
465
466 In previous studies of mixed-length ligand layers with full coverage,
467 we observed that ligand-solvent alignment was an important factor for
468 heat transfer into the solvent. With high surface curvature and lower
469 effective coverages, ligand behavior also becomes more complex. Some
470 chains may be lying down on the surface, and solvent may not be
471 penetrating the ligand layer to the same degree as in the planar
472 surfaces.
473
474 Additionally, the ligand flexibility directly alters the vibrational
475 density of states for the layer that mediates the transfer of phonons
476 between the metal and the solvent. This could be a partial explanation
477 for the observed differences between the fully conjugated and more
478 flexible ligands.
479
480 In the following sections we provide details on how we
481 measure surface corrugation, solvent-ligand interpenetration, and
482 ordering of the solvent and ligand at the surfaces of the
483 nanospheres. We also investigate the overlap between vibrational
484 densities of states for the various ligands.
485
486 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487 % CORRUGATION OF PARTICLE SURFACE
488 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
489 \subsection{Corrugation of the Particle Surface}
490
491 The bonding sites for thiols on gold surfaces have been studied
492 extensively and include configurations beyond the traditional atop,
493 bridge, and hollow sites found on planar surfaces. In particular, the
494 deep potential well between the gold atoms and the thiolate sulfur
495 atoms leads to insertion of the sulfur into the gold lattice and
496 displacement of interfacial gold atoms. The degree of ligand-induced
497 surface restructuring may have an impact on the interfacial thermal
498 conductance and is an important phenomenon to quantify.
499
500 Henz, \textit{et al.}\cite{Henz:2008qf} used the metal
501 density as a function of radius to measure the degree of mixing
502 between the thiol sulfurs and surface gold atoms at the edge of a
503 nanoparticle. Although metal density is important, disruption of the
504 local crystalline ordering would also have a large effect on the
505 phonon spectrum in the particles. To measure this effect, we use the
506 fraction of gold atoms exhibiting local fcc ordering as a function of
507 radius to describe the ligand-induced disruption of the nanoparticle
508 surface.
509
510 The local bond orientational order can be described using the method
511 of Steinhardt \textit{et al.}\cite{Steinhardt1983} The local bonding
512 environment, $\bar{q}_{\ell m}$, for each atom in the system is
513 determined by averaging over the spherical harmonics between that atom
514 and each of its neighbors,
515 \begin{equation}
516 \bar{q}_{\ell m} = \sum_i Y_\ell^m(\theta_i, \phi_i)
517 \end{equation}
518 where $\theta_i$ and $\phi_i$ are the relative angular coordinates of
519 neighbor $i$ in the laboratory frame. A global average orientational
520 bond order parameter, $\bar{Q}_{\ell m}$, is the average over each
521 $\bar{q}_{\ell m}$ for all atoms in the system. To remove the
522 dependence on the laboratory coordinate frame, the third order
523 rotationally invariant combination of $\bar{Q}_{\ell m}$,
524 $\hat{w}_\ell$, is utilized here.\cite{Steinhardt1983,Vardeman:2008fk}
525
526 For $\ell=4$, the ideal face-centered cubic (fcc), body-centered cubic
527 (bcc), hexagonally close-packed (hcp), and simple cubic (sc) local
528 structures exhibit $\hat{w}_4$ values of -0.159, 0.134, 0.159, and
529 0.159, respectively. Because $\hat{w}_4$ exhibits an extreme value for
530 fcc structures, it is ideal for measuring local fcc
531 ordering. The spatial distribution of $\hat{w}_4$ local bond
532 orientational order parameters, $p(\hat{w}_4 , r)$, can provide
533 information about the location of individual atoms that are central to
534 local fcc ordering.
535
536 The fraction of fcc-ordered gold atoms at a given radius in the
537 nanoparticle,
538 \begin{equation}
539 f_\mathrm{fcc}(r) = \int_{-\infty}^{w_c} p(\hat{w}_4, r) d \hat{w}_4
540 \end{equation}
541 is described by the distribution of the local bond orientational order
542 parameters, $p(\hat{w}_4, r)$, and $w_c$, a cutoff for the peak
543 $\hat{w}_4$ value displayed by fcc structures. A $w_c$ value of -0.12
544 was chosen to isolate the fcc peak in $\hat{w}_4$.
545
546 As illustrated in Figure \ref{fig:Corrugation}, the presence of
547 ligands decreases the fcc ordering of the gold atoms at the
548 nanoparticle surface. For the smaller nanoparticles, this disruption
549 extends into the core of the nanoparticle, indicating widespread
550 disruption of the lattice.
551
552 \begin{figure}
553 \includegraphics[width=\linewidth]{figures/fcc}
554 \caption{Fraction of gold atoms with fcc ordering as a function of
555 radius for a 10 \AA\ radius nanoparticle. The decreased fraction
556 of fcc-ordered atoms in ligand-protected nanoparticles relative to
557 bare particles indicates restructuring of the nanoparticle surface
558 by the thiolate sulfur atoms.}
559 \label{fig:Corrugation}
560 \end{figure}
561
562 We may describe the thickness of the disrupted nanoparticle surface by
563 defining a corrugation factor, $c$, as the ratio of the radius at
564 which the fraction of gold atoms with fcc ordering is 0.9 and the
565 radius at which the fraction is 0.5.
566
567 \begin{equation}
568 c = 1 - \frac{r(f_\mathrm{fcc} = 0.9)}{r(f_\mathrm{fcc} = 0.5)}
569 \end{equation}
570
571 A sharp interface will have a steep drop in $f_\mathrm{fcc}$ at the
572 edge of the particle ($c \rightarrow$ 0). In the opposite limit where
573 the entire nanoparticle surface is restructured by ligands, the radius
574 at which there is a high probability of fcc ordering moves
575 dramatically inward ($c \rightarrow$ 1).
576
577 The computed corrugation factors are shown in Figure
578 \ref{fig:NPthiols_corrugation} for bare nanoparticles and for
579 ligand-protected particles as a function of ligand chain length. The
580 largest nanoparticles are only slightly restructured by the presence
581 of ligands on the surface, while the smallest particle ($r$ = 10 \AA)
582 exhibits significant disruption of the original fcc ordering when
583 covered with a half-monolayer of thiol ligands.
584
585 \begin{figure}
586 \includegraphics[width=\linewidth]{figures/C3.pdf}
587 \caption{Computed corrugation values for 4 sizes of solvated
588 nanoparticles that are bare or protected with a 50\% coverage of
589 C$_{4}$, C$_{8}$, or C$_{12}$ thiolate ligands. The smallest (10
590 \AA ) particles show significant disruption to their crystal
591 structures, and the length and stiffness of the ligands is a
592 contributing factor to the surface disruption.}
593 \label{fig:NPthiols_corrugation}
594 \end{figure}
595
596 Because the thiolate ligands do not significantly alter the larger
597 particle crystallinity, the surface corrugation does not seem to be a
598 likely candidate to explain the large increase in thermal conductance
599 at the interface when ligands are added.
600
601 % \begin{equation}
602 % C = \frac{r_{bare}(\rho_{\scriptscriptstyle{0.85}}) - r_{capped}(\rho_{\scriptscriptstyle{0.85}})}{r_{bare}(\rho_{\scriptscriptstyle{0.85}})}.
603 % \end{equation}
604 %
605 % Here, $r_{bare}(\rho_{\scriptscriptstyle{0.85}})$ is the radius of a bare nanoparticle at which the density is $85\%$ the bulk value and $r_{capped}(\rho_{\scriptscriptstyle{0.85}})$ is the corresponding radius for a particle of the same size with a layer of ligands. $C$ has a value of 0 for a bare particle and approaches $1$ as the degree of surface atom mixing increases.
606
607
608
609
610 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
611 % MOBILITY OF INTERFACIAL SOLVENT
612 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
613 % \subsection{Mobility of Interfacial Solvent}
614
615 % Another possible mechanism for increasing interfacial conductance is
616 % the mobility of the interfacial solvent. We used a survival
617 % correlation function, $C(t)$, to measure the residence time of a
618 % solvent molecule in the nanoparticle thiolate
619 % layer.\cite{Stocker:2013cl} This function correlates the identity of
620 % all hexane molecules within the radial range of the thiolate layer at
621 % two separate times. If the solvent molecule is present at both times,
622 % the configuration contributes a $1$, while the absence of the molecule
623 % at the later time indicates that the solvent molecule has migrated
624 % into the bulk, and this configuration contributes a $0$. A steep decay
625 % in $C(t)$ indicates a high turnover rate of solvent molecules from the
626 % chain region to the bulk. We may define the escape rate for trapped
627 % solvent molecules at the interface as
628 % \begin{equation}
629 % k_\mathrm{escape} = \left( \int_0^T C(t) dt \right)^{-1}
630 % \label{eq:mobility}
631 % \end{equation}
632 % where T is the length of the simulation. This is a direct measure of
633 % the rate at which solvent molecules initially entangled in the
634 % thiolate layer can escape into the bulk. When $k_\mathrm{escape}
635 % \rightarrow 0$, the solvent becomes permanently trapped in the
636 % interfacial region.
637
638 % The solvent escape rates for bare and ligand-protected nanoparticles
639 % are shown in Figure \ref{fig:NPthiols_combo}. As the ligand chain
640 % becomes longer and more flexible, interfacial solvent molecules become
641 % trapped in the ligand layer and the solvent escape rate decreases.
642 % This mechanism contributes a partial explanation as to why the longer
643 % ligands have significantly lower thermal conductance.
644
645 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
646 % ORIENTATION OF LIGAND CHAINS
647 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
648 \subsection{Orientation of Ligand Chains}
649
650 As the saturated ligand chain length increases in length, it exhibits
651 significantly more conformational flexibility. Thus, different lengths
652 of ligands should favor different chain orientations on the surface of
653 the nanoparticle. To determine the distribution of ligand orientations
654 relative to the particle surface we examine the probability of finding
655 a ligand with a particular orientation relative to the surface normal
656 of the nanoparticle,
657 \begin{equation}
658 \cos{(\theta)}=\frac{\vec{r}_i\cdot\hat{u}_i}{|\vec{r}_i||\hat{u}_i|}
659 \end{equation}
660 where $\vec{r}_{i}$ is the vector between the cluster center of mass
661 and the sulfur atom on ligand molecule {\it i}, and $\hat{u}_{i}$ is
662 the vector between the sulfur atom and \ce{CH3} pseudo-atom on ligand
663 molecule {\it i}. As depicted in Figure \ref{fig:NP_pAngle}, $\theta
664 \rightarrow 180^{\circ}$ for a ligand chain standing upright on the
665 particle ($\cos{(\theta)} \rightarrow -1$) and $\theta \rightarrow
666 90^{\circ}$ for a ligand chain lying down on the surface
667 ($\cos{(\theta)} \rightarrow 0$). As the thiolate alkane chain
668 increases in length and becomes more flexible, the ligands are more
669 willing to lie down on the nanoparticle surface and exhibit increased
670 population at $\cos{(\theta)} = 0$.
671
672 \begin{figure}
673 \includegraphics[width=\linewidth]{figures/NP_pAngle}
674 \caption{The two extreme cases of ligand orientation relative to the
675 nanoparticle surface: the ligand completely outstretched
676 ($\cos{(\theta)} = -1$) and the ligand fully lying down on the
677 particle surface ($\cos{(\theta)} = 0$).}
678 \label{fig:NP_pAngle}
679 \end{figure}
680
681 An order parameter describing the average ligand chain orientation relative to
682 the nanoparticle surface is available using the second order Legendre
683 parameter,
684 \begin{equation}
685 P_2 = \left< \frac{1}{2} \left(3\cos^2(\theta) - 1 \right) \right>
686 \end{equation}
687
688 Ligand populations that are perpendicular to the particle surface have
689 $P_2$ values of 1, while ligand populations lying flat on the
690 nanoparticle surface have $P_2$ values of $-0.5$. Disordered ligand
691 layers will exhibit mean $P_2$ values of 0. As shown in Figure
692 \ref{fig:NPthiols_P2} the ligand $P_2$ values approaches 0 as
693 ligand chain length -- and ligand flexibility -- increases.
694
695 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
696 % ORIENTATION OF INTERFACIAL SOLVENT
697 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
698 \subsection{Orientation of Interfacial Solvent}
699
700 Similarly, we examined the distribution of \emph{hexane} molecule
701 orientations relative to the particle surface using the same angular
702 analysis utilized for the ligand chain orientations. In this case,
703 $\vec{r}_i$ is the vector between the particle center of mass and one
704 of the \ce{CH2} pseudo-atoms in the middle of hexane molecule $i$ and
705 $\hat{u}_i$ is the vector between the two \ce{CH3} pseudo-atoms on
706 molecule $i$. Since we are only interested in the orientation of
707 solvent molecules near the ligand layer, we select only the hexane
708 molecules within a specific $r$-range, between the edge of the
709 particle and the end of the ligand chains. A large population of
710 hexane molecules with $\cos{(\theta)} \sim \pm 1$ would indicate
711 interdigitation of the solvent molecules between the upright ligand
712 chains. A more random distribution of $\cos{(\theta)}$ values
713 indicates a disordered arrangement of solvent molecules near the particle
714 surface. Again, $P_2$ order parameter values provide a population
715 analysis for the solvent that is close to the particle surface.
716
717 The average orientation of the interfacial solvent molecules is
718 notably flat on the \emph{bare} nanoparticle surfaces. This blanket of
719 hexane molecules on the particle surface may act as an insulating
720 layer, increasing the interfacial thermal resistance. As the length
721 (and flexibility) of the ligand increases, the average interfacial
722 solvent P$_2$ value approaches 0, indicating a more random orientation
723 of the ligand chains. The average orientation of solvent within the
724 $C_8$ and $C_{12}$ ligand layers is essentially random. Solvent
725 molecules in the interfacial region of $C_4$ ligand-protected
726 nanoparticles do not lie as flat on the surface as in the case of the
727 bare particles, but are not as randomly oriented as the longer ligand
728 lengths.
729
730 \begin{figure}
731 \includegraphics[width=\linewidth]{figures/P2_3.pdf}
732 \caption{Computed ligand and interfacial solvent orientational $P_2$
733 values for 4 sizes of solvated nanoparticles that are bare or
734 protected with a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$
735 alkanethiolate ligands. Increasing stiffness of the ligand orients
736 these molecules normal to the particle surface, while the length
737 of the ligand chains works to prevent solvent from lying flat on
738 the surface.}
739 \label{fig:NPthiols_P2}
740 \end{figure}
741
742 These results are particularly interesting in light of our previous
743 results\cite{Stocker:2013cl}, where solvent molecules readily filled
744 the vertical gaps between neighboring ligand chains and there was a
745 strong correlation between ligand and solvent molecular
746 orientations. It appears that the introduction of surface curvature
747 and a lower ligand packing density creates a disordered ligand layer
748 that lacks well-formed channels for the solvent molecules to occupy.
749
750 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
751 % SOLVENT PENETRATION OF LIGAND LAYER
752 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
753 \subsection{Solvent Penetration of Ligand Layer}
754
755 The extent of ligand -- solvent interaction is also determined by the
756 degree to which these components occupy the same region of space
757 adjacent to the nanoparticle. The radial density profiles of these
758 components help determine this degree of interaction. Figure
759 \ref{fig:density} shows representative density profiles for solvated
760 25 \AA\ radius nanoparticles with no ligands, and with a 50\% coverage
761 of C$_{4}$, C$_{8}$, and C$_{12}$ thiolates.
762
763 \begin{figure}
764 \includegraphics[width=\linewidth]{figures/density}
765 \caption{Radial density profiles for 25 \AA\ radius nanoparticles
766 with no ligands (circles), C$_{4}$ ligands (squares), C$_{8}$
767 ligands (diamonds), and C$_{12}$ ligands (triangles). Ligand
768 density is indicated with filled symbols, solvent (hexane) density
769 is indicated with open symbols. As ligand chain length increases,
770 the nearby solvent is excluded from the ligand layer. The
771 conjugated ligands (upper panel) can create a separated solvent
772 shell within the ligand layer and also allow significantly more
773 solvent to penetrate close to the particle.}
774 \label{fig:density}
775 \end{figure}
776
777 The differences between the radii at which the hexane surrounding the
778 ligand-covered particles reaches bulk density correspond nearly
779 exactly to the differences between the lengths of the ligand
780 chains. Beyond the edge of the ligand layer, the solvent reaches its
781 bulk density within a few angstroms. The differing shapes of the
782 density curves indicate that the solvent is increasingly excluded from
783 the ligand layer as the chain length increases.
784
785 The conjugated ligands create a distinct solvent shell within the
786 ligand layer and also allow significantly more solvent to penetrate
787 close to the particle. We define a density overlap parameter,
788 \begin{equation}
789 O_{l-s} = \frac{1}{V} \int_0^{r_\mathrm{max}} 4 \pi r^2 \frac{4 \rho_l(r) \rho_s(r)}{\left(\rho_l(r) +
790 \rho_s(r)\right)^2} dr
791 \end{equation}
792 where $\rho_l(r)$ and $\rho_s(r)$ are the ligand and solvent densities
793 at a radius $r$, and $V$ is the total integration volume
794 ($V = 4\pi r_\mathrm{max}^3 / 3$). The fraction in the integrand is a
795 dimensionless quantity that is unity when ligand and solvent densities
796 are identical at radius $r$, but falls to zero when either of the two
797 components are excluded from that region.
798
799 \begin{figure}
800 \includegraphics[width=\linewidth]{figures/rho3}
801 \caption{Density overlap parameters ($O_{l-s}$) for solvated
802 nanoparticles protected by thiolate ligands. In general, the
803 rigidity of the fully-conjugated ligands provides the easiest
804 route for solvent to enter the interfacial region. Additionally,
805 shorter chains allow a greater degree of solvent penetration of
806 the ligand layer.}
807 \label{fig:rho3}
808 \end{figure}
809
810 The density overlap parameters are shown in Fig. \ref{fig:rho3}. The
811 calculated overlap parameters indicate that the conjugated ligand
812 allows for the most solvent penetration close to the particle, and
813 that shorter chains generally permit greater solvent penetration in
814 the interfacial region. Increasing overlap can certainly allow for
815 enhanced thermal transport, but this is clearly not the only
816 contributing factor. Even when the solvent and ligand are in close
817 physical contact, there must also be good vibrational overlap between
818 the phonon densities of states in the ligand and solvent to transmit
819 vibrational energy between the two materials.
820
821 \subsection{Ligand-mediated Vibrational Overlap}
822
823 In phonon scattering models for interfacial thermal
824 conductance,\cite{Swartz:1989uq,Young:1989xy,Cahill:2003fk,Reddy:2005fk,Schmidt:2010nr}
825 the frequency-dependent transmission probability
826 ($t_{a \rightarrow b}(\omega)$) predicts phonon transfer between
827 materials $a$ and $b$. Many of the models for interfacial phonon
828 transmission estimate this quantity using the phonon density of states
829 and group velocity, and make use of a Debye model for the density of
830 states in the solid.
831
832 A consensus picture is that in order to transfer the energy carried by
833 an incoming phonon of frequency $\omega$ on the $a$ side, the phonon
834 density of states on the $b$ side must have a phonon of the same
835 frequency. The overlap of the phonon densities of states, particularly
836 at low frequencies, therefore contributes to the transfer of heat.
837 Phonon scattering must also be done in a direction perpendicular to
838 the interface. In the geometries described here, there are two
839 interfaces (particle $\rightarrow$ ligand, and ligand $\rightarrow$
840 solvent), and the vibrational overlap between the ligand and the other
841 two components is going to be relevant to heat transfer.
842
843 To estimate the relevant densities of states, we have projected the
844 velocity of each atom $i$ in the region of the interface onto a
845 direction normal to the interface. For the nanosphere geometries
846 studied here, the normal direction depends on the instantaneous
847 positon of the atom relative to the center of mass of the particle.
848 \begin{equation}
849 v_\perp(t) = \mathbf{v}(t) \cdot \frac{\mathbf{r}(t)}{\left|\mathbf{r}(t)\right|}
850 \end{equation}
851 The quantity $v_\perp(t)$ measures the instantaneous velocity of an
852 atom in a direction perpendicular to the nanoparticle interface. In
853 the interfacial region, the autocorrelation function of these
854 velocities,
855 \begin{equation}
856 C_\perp(t) = \left< v_\perp(t) \cdot v_\perp(0) \right>,
857 \end{equation}
858 will include contributions from all of the phonon modes present at the
859 interface. The Fourier transform of the time-symmetrized
860 autocorrelation function provides an estimate of the vibrational
861 density of states,\cite{Shin:2010sf}
862 \begin{equation}
863 \rho(\omega) = \frac{1}{\tau} \int_{-\tau/2}^{\tau/2} C_\perp(t) e^{-i
864 \omega t} dt.
865 \end{equation}
866 Here $\tau$ is the total observation time for the autocorrelation
867 function. In Fig.~\ref{fig:vdos} we show the low-frequency region of
868 the normalized vibrational densities of states for the three chemical
869 components (gold nanoparticle, C$_{12}$ ligands, and interfacial
870 solvent). The double bond in the penultimate location is a small
871 perturbation on ligands of this size, and that is reflected in
872 relatively similar spectra in the lower panels. The fully conjugated
873 ligand, however, pushes the peak in the lowest frequency band from
874 $\sim 29 \mathrm{cm}^{-1}$ to $\sim 55 \mathrm{cm}^{-1}$, yielding
875 significant overlap with the density of states in the nanoparticle.
876 This ligand also increases the overlap with the solvent density of
877 states in a band between 280 and 380 $\mathrm{cm}^{-1}$. This
878 provides some physical basis for the high interfacial conductance
879 observed for the fully conjugated $C_8$ and $C_{12}$ ligands.
880
881 \begin{figure}
882 \includegraphics[width=\linewidth]{figures/rho_omega_12}
883 \caption{The low frequency portion of the vibrational density of
884 states for three chemical components (gold nanoparticles, C$_{12}$
885 ligands, and hexane solvent). These densities of states were
886 computed using the velocity autocorrelation functions for atoms in
887 the interfacial region, constructed with velocities projected onto
888 a direction normal to the interface.}
889 \label{fig:vdos}
890 \end{figure}
891
892 The similarity between the density of states for the alkanethiolate
893 and penultimate ligands also helps explain why the interfacial
894 conductance is nearly the same for these two ligands, particularly at
895 longer chain lengths.
896
897 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
898 % DISCUSSION
899 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
900 \section{Discussion}
901
902 The chemical bond between the metal and the ligand introduces
903 vibrational overlap that is not present between the bare metal surface
904 and solvent. Thus, regardless of ligand identity or chain length, the
905 presence of a half-monolayer ligand coverage yields a higher
906 interfacial thermal conductance value than the bare nanoparticle. The
907 mechanism for the varying conductance for the different ligands is
908 somewhat less clear. Ligand-based alterations to vibrational density
909 of states is a major contributor, but some of the ligands can disrupt
910 the crystalline structure of the smaller nanospheres, while others can
911 re-order the interfacial solvent and alter the interpenetration
912 profile between ligand and solvent chains. Further work to separate
913 the effects of ligand-solvent interpenetration and surface
914 reconstruction is clearly needed for a complete picture of the heat
915 transport in these systems.
916
917 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
918 % **ACKNOWLEDGMENTS**
919 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
920 \begin{acknowledgments}
921 Support for this project was provided by the National Science Foundation
922 under grant CHE-1362211. Computational time was provided by the
923 Center for Research Computing (CRC) at the University of Notre Dame.
924 \end{acknowledgments}
925
926 \newpage
927 \bibliographystyle{aip}
928 \bibliography{NPthiols}
929
930 \end{document}

Properties

Name Value
svn:executable