ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/NPthiols/NPthiols.tex
Revision: 4161
Committed: Thu May 29 16:33:41 2014 UTC (10 years, 1 month ago) by kstocke1
Content type: application/x-tex
File size: 37661 byte(s)
Log Message:
Final edits and submission zip

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3
4 \usepackage{caption}
5 \usepackage{geometry}
6 \usepackage{natbib}
7 \usepackage{setspace}
8 \usepackage{xkeyval}
9 %%%%%%%%%%%%%%%%%%%%%%%
10 \usepackage{amsmath}
11 \usepackage{amssymb}
12 \usepackage{times}
13 \usepackage{mathptm}
14 \usepackage{caption}
15 \usepackage{tabularx}
16 \usepackage{longtable}
17 \usepackage{graphicx}
18 \usepackage{achemso}
19 \usepackage{wrapfig}
20 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
21 \usepackage{url}
22
23 \title{Interfacial Thermal Conductance of Alkanethiolate-Protected Gold
24 Nanospheres}
25
26 \author{Kelsey M. Stocker}
27 \author{J. Daniel Gezelter}
28 \email{gezelter@nd.edu}
29 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
30 Department of Chemistry and Biochemistry\\
31 University of Notre Dame\\
32 Notre Dame, Indiana 46556}
33
34
35 \keywords{Nanoparticles, interfaces, thermal conductance}
36
37 \begin{document}
38
39 \begin{tocentry}
40 \center\includegraphics[width=3.9cm]{figures/TOC}
41 \end{tocentry}
42
43 \newcolumntype{A}{p{1.5in}}
44 \newcolumntype{B}{p{0.75in}}
45
46
47 \begin{abstract}
48 Molecular dynamics simulations of alkanethiolate-protected and
49 solvated gold nanoparticles were carried out in the presence of a
50 non-equilibrium heat flux between the solvent and the core of the
51 particle. The interfacial thermal conductance ($G$) was computed for
52 these interfaces, and the behavior of the thermal conductance was
53 studied as a function of particle size and ligand chain length. In
54 all cases, thermal conductance of the ligand-protected particles was
55 higher than the bare metal--solvent interface. A number of
56 mechanisms for the enhanced conductance were investigated, including
57 thiolate-driven corrugation of the metal surface, solvent mobility
58 and ordering at the interface, and ligand ordering relative to the
59 particle surface. The shortest and least flexible ligand, butanethiolate,
60 exhibited the highest interfacial thermal conductance and was the
61 least likely to trap solvent molecules within the ligand layer. At
62 the 50\% coverage levels studied, heat transfer into the solvent
63 relies primarily on convective motion of the solvent molecules from
64 the surface of the particle into the bulk. This mode of heat
65 transfer is reduced by slow solvent escape rates, and this effect was
66 observed to lower the interfacial conductance for the longer-chain ligands.
67 \end{abstract}
68
69 \newpage
70
71 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72 % INTRODUCTION
73 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74 \section{Introduction}
75
76 Heat transport across various nanostructured interfaces has been
77 the subject of intense experimental
78 interest,\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101,Wang10082007,doi:10.1021/jp8051888,PhysRevB.80.195406,doi:10.1021/la904855s}
79 and the interfacial thermal conductance, $G$, is the principal quantity of
80 interest for understanding interfacial heat
81 transport.\cite{cahill:793} Because nanoparticles have a significant
82 fraction of their atoms at the particle / solvent interface, the
83 chemical details of these interfaces govern the thermal transport
84 properties.
85
86 Previously, reverse nonequilibrium molecular dynamics (RNEMD) methods
87 have been applied to calculate the interfacial thermal conductance at
88 flat (111) metal / organic solvent interfaces that had been chemically
89 protected by mixed-chain alkanethiolate groups.\cite{kuang:AuThl}
90 These simulations suggested an explanation for the increased thermal
91 conductivity at alkanethiol-capped metal surfaces compared with bare
92 metal interfaces. Specifically, the chemical bond between the metal
93 and the ligand introduces a vibrational overlap that is not present
94 without the protecting group, and the overlap between the vibrational
95 spectra (metal to ligand, ligand to solvent) provides a mechanism for
96 rapid thermal transport across the interface. The simulations also
97 suggested that this phenomenon is a non-monotonic function of the
98 fractional coverage of the surface, as moderate coverages allow
99 diffusive heat transport of solvent molecules that come into close
100 contact with the ligands.
101
102 Simulations of {\it mixed-chain} alkylthiolate surfaces showed that
103 solvent trapped close to the interface can be efficient at moving
104 thermal energy away from the surface.\cite{Stocker:2013cl} Trapped
105 solvent molecules that were aligned with nearby
106 ligands (but which were less able to diffuse into the bulk) were able
107 to increase the thermal conductance of the interface. This indicates
108 that the ligand-to-solvent vibrational energy transfer is a key
109 feature for increasing particle-to-solvent thermal conductance.
110
111 Recently, we extended RNEMD methods for use in non-periodic geometries
112 by creating scaling/shearing moves between concentric regions of a
113 simulation.\cite{Stocker:2014qq} In this work, we apply this
114 non-periodic variant of RNEMD to investigate the role that {\it
115 curved} nanoparticle surfaces play in heat and mass transport. On
116 planar surfaces, we discovered that orientational ordering of surface
117 protecting ligands had a large effect on the heat conduction from the
118 metal to the solvent. Smaller nanoparticles have high surface
119 curvature that creates gaps in well-ordered self-assembled monolayers,
120 and the effect of those gaps on the thermal conductance is unknown.
121
122
123
124 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
125 % INTERFACIAL THERMAL CONDUCTANCE OF METALLIC NANOPARTICLES
126 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
127 %\section{Interfacial Thermal Conductance of Metallic Nanoparticles}
128
129 For a solvated nanoparticle, it is possible to define a critical value
130 for the interfacial thermal conductance,
131 \begin{equation}
132 G_c = \frac{3 C_s \Lambda_s}{R C_p}
133 \end{equation}
134 which depends on the solvent heat capacity, $C_s$, solvent thermal
135 conductivity, $\Lambda_s$, particle radius, $R$, and nanoparticle heat
136 capacity, $C_p$.\cite{Wilson:2002uq} In the limit of infinite
137 interfacial thermal conductance, $G \gg G_c$, cooling of the
138 nanoparticle is limited by the solvent properties, $C_s$ and
139 $\Lambda_s$. In the opposite limit, $G \ll G_c$, the heat dissipation
140 is controlled by the thermal conductance of the particle / fluid
141 interface. It is this regime with which we are concerned, where
142 properties of ligands and the particle surface may be tuned to
143 manipulate the rate of cooling for solvated nanoparticles. Based on
144 estimates of $G$ from previous simulations as well as experimental
145 results for solvated nanostructures, gold nanoparticles solvated in
146 hexane are in the $G \ll G_c$ regime for radii smaller than 40 nm. The
147 particles included in this study are more than an order of magnitude
148 smaller than this critical radius, so the heat dissipation should be
149 controlled entirely by the surface features of the particle / ligand /
150 solvent interface.
151
152 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153 % STRUCTURE OF SELF-ASSEMBLED MONOLAYERS ON NANOPARTICLES
154 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
155 \subsection{Structures of Self-Assembled Monolayers on Nanoparticles}
156
157 Though the ligand packing on planar surfaces has been characterized for many
158 different ligands and surface facets, it is not obvious \emph{a
159 priori} how the same ligands will behave on the highly curved
160 surfaces of spherical nanoparticles. Thus, as new applications of
161 ligand-stabilized nanostructures have been proposed, the structure
162 and dynamics of ligands on metallic nanoparticles have been studied
163 using molecular simulation,\cite{Henz2007,Henz:2008qf} NMR, XPS, FTIR, calorimetry, and surface microscopies.\cite{Badia1996:2,Badia1996,Badia1997:2,Badia1997,Badia2000}
164 Badia, \textit{et al.} used transmission electron microscopy to
165 determine that alkanethiol ligands on gold nanoparticles pack
166 approximately 30\% more densely than on planar Au(111)
167 surfaces.\cite{Badia1996:2} Subsequent experiments demonstrated that
168 even at full coverages, surface curvature creates voids between linear
169 ligand chains that can be filled via interdigitation of ligands on
170 neighboring particles.\cite{Badia1996} The molecular dynamics
171 simulations of Henz, \textit{et al.} indicate that at low coverages,
172 the thiolate alkane chains will lie flat on the nanoparticle
173 surface\cite{Henz2007,Henz:2008qf} Above 90\% coverage, the ligands stand upright
174 and recover the rigidity and tilt angle displayed on planar
175 facets. Their simulations also indicate a high degree of mixing
176 between the thiolate sulfur atoms and surface gold atoms at high
177 coverages.
178
179 In this work, thiolated gold nanospheres were modeled using a united atom force field and non-equilibrium molecular dynamics. Gold nanoparticles
180 with radii ranging from 10 - 25 \AA\ were created from a bulk fcc
181 lattice. To match surface coverages previously reported by Badia,
182 \textit{et al.}\cite{Badia1996:2}, these particles were passivated
183 with a 50\% coverage of a selection of alkyl thiolates of varying
184 chain lengths. The passivated particles were then solvated in hexane.
185 Details of the models and simulation protocol follow in the next
186 section.
187
188 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189 % NON-PERIODIC VSS-RNEMD METHODOLOGY
190 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191 \subsection{Creating a thermal flux between particles and solvent}
192
193 The non-periodic variant of VSS-RNEMD\cite{Stocker:2014qq} applies a
194 series of velocity scaling and shearing moves at regular intervals to
195 impose a flux between two concentric spherical regions. To impose a
196 thermal flux between the shells (without an accompanying angular
197 shear), we solve for scaling coefficients $a$ and $b$,
198 \begin{eqnarray}
199 a = \sqrt{1 - \frac{q_r \Delta t}{K_a - K_a^\mathrm{rot}}}\\ \nonumber\\
200 b = \sqrt{1 + \frac{q_r \Delta t}{K_b - K_b^\mathrm{rot}}}
201 \end{eqnarray}
202 at each time interval. These scaling coefficients conserve total
203 kinetic energy and angular momentum subject to an imposed heat rate,
204 $q_r$. The coefficients also depend on the instantaneous kinetic
205 energy, $K_{\{a,b\}}$, and the total rotational kinetic energy of each
206 shell, $K_{\{a,b\}}^\mathrm{rot} = \sum_i m_i \left( \mathbf{v}_i
207 \times \mathbf{r}_i \right)^2 / 2$.
208
209 The scaling coefficients are determined and the velocity changes are
210 applied at regular intervals,
211 \begin{eqnarray}
212 \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~~\:\\
213 \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.
214 \end{eqnarray}
215 Here $\left < \omega_a \right > \times \mathbf{r}_i$ is the
216 contribution to the velocity of particle $i$ due to the overall
217 angular velocity of the $a$ shell. In the absence of an angular
218 momentum flux, the angular velocity $\left < \omega_a \right >$ of the
219 shell is nearly 0 and the resultant particle velocity is a nearly
220 linear scaling of the initial velocity by the coefficient $a$ or $b$.
221
222 Repeated application of this thermal energy exchange yields a radial
223 temperature profile for the solvated nanoparticles that depends
224 linearly on the applied heat rate, $q_r$. Similar to the behavior in
225 the slab geometries, the temperature profiles have discontinuities at
226 the interfaces between dissimilar materials. The size of the
227 discontinuity depends on the interfacial thermal conductance, which is
228 the primary quantity of interest.
229
230 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
231 % CALCULATING TRANSPORT PROPERTIES
232 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
233 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
234 % INTERFACIAL THERMAL CONDUCTANCE
235 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
236 \subsection{Interfacial Thermal Conductance}
237
238 As described in earlier work,\cite{Stocker:2014qq} the thermal
239 conductance of each spherical shell may be defined as the inverse
240 Kapitza resistance of the shell. To describe the thermal conductance
241 of an interface of considerable thickness -- such as the ligand layers
242 shown here -- we can sum the individual thermal resistances of each
243 concentric spherical shell to arrive at the inverse of the total
244 interfacial thermal conductance. In slab geometries, the intermediate
245 temperatures cancel, but for concentric spherical shells, the
246 intermediate temperatures and surface areas remain in the final sum,
247 requiring the use of a series of individual resistance terms:
248
249 \begin{equation}
250 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
251 T_i\right) 4 \pi r_i^2.
252 \end{equation}
253
254 The longest ligand considered here is in excess of 15 \AA\ in length,
255 and we use 10 concentric spherical shells to describe the total
256 interfacial thermal conductance of the ligand layer.
257
258 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
259 % COMPUTATIONAL DETAILS
260 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
261 \section{Computational Details}
262
263 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264 % FORCE FIELDS
265 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
266 \subsection{Force Fields}
267
268 Throughout this work, gold -- gold interactions are described by the
269 quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} The hexane
270 solvent is described by the TraPPE united atom
271 model,\cite{TraPPE-UA.alkanes} where sites are located at the carbon
272 centers for alkyl groups. The TraPPE-UA model for hexane provides both
273 computational efficiency and reasonable accuracy for bulk thermal
274 conductivity values. Bonding interactions were used for
275 intra-molecular sites closer than 3 bonds. Effective Lennard-Jones
276 potentials were used for non-bonded interactions.
277
278 To describe the interactions between metal (Au) and non-metal atoms,
279 potential energy terms were adapted from an adsorption study of alkyl
280 thiols on gold surfaces by Vlugt, \textit{et
281 al.}\cite{vlugt:cpc2007154} They fit an effective pair-wise
282 Lennard-Jones form of potential parameters for the interaction between
283 Au and pseudo-atoms CH$_x$ and S based on a well-established and
284 widely-used effective potential of Hautman and Klein for the Au(111)
285 surface.\cite{hautman:4994}
286
287
288
289 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
290 % SIMULATION PROTOCOL
291 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
292 \subsection{Simulation Protocol}
293
294 Gold nanospheres with radii ranging from 10 - 25 \AA\ were created
295 from a bulk fcc lattice and were thermally equilibrated prior to the
296 addition of ligands. A 50\% coverage of ligands (based on coverages
297 reported by Badia, \textit{et al.}\cite{Badia1996:2}) were placed on
298 the surface of the equilibrated nanoparticles using
299 Packmol\cite{packmol}. We have chosen three lengths of ligands: butanethiolate ($C_4$), octanethiolate ($C_8$), and dodecanethiolate ($C_{12}$). The nanoparticle / ligand complexes were
300 thermally equilibrated to allow for ligand conformational flexibility. Packmol was then used to solvate the
301 structures inside a spherical droplet of hexane. The thickness of the
302 solvent layer was chosen to be at least 1.5$\times$ the combined
303 radius of the nanoparticle / ligand structure. The fully solvated
304 system was equilibrated for at least 1 ns using the Langevin Hull to
305 apply 50 atm of pressure and a target temperature of 250
306 K.\cite{Vardeman2011} Typical system sizes ranged from 18,310 united
307 atom sites for the 10 \AA\ particles with $C_4$ ligands to 89,490 sites
308 for the 25 \AA\ particles with $C_{12}$ ligands. Figure
309 \ref{fig:NP25_C12h1} shows one of the solvated 25 \AA\ nanoparticles
310 passivated with the $C_{12}$ ligands.
311
312 \begin{figure}
313 \includegraphics[width=\linewidth]{figures/NP25_C12h1}
314 \caption{A 25 \AA\ radius gold nanoparticle protected with a
315 half-monolayer of TraPPE-UA dodecanethiolate (C$_{12}$)
316 ligands and solvated in TraPPE-UA hexane. The interfacial
317 thermal conductance is computed by applying a kinetic energy
318 flux between the nanoparticle and an outer shell of
319 solvent.}
320 \label{fig:NP25_C12h1}
321 \end{figure}
322
323 Once equilibrated, thermal fluxes were applied for 1 ns, until stable
324 temperature gradients had developed. Systems were run under moderate
325 pressure (50 atm) with an average temperature (250K) that maintained a
326 compact solvent cluster and avoided formation of a vapor layer near
327 the heated metal surface. Pressure was applied to the system via the
328 non-periodic Langevin Hull.\cite{Vardeman2011} However, thermal
329 coupling to the external temperature bath was removed to avoid
330 interference with the imposed RNEMD flux.
331
332 Because the method conserves \emph{total} angular momentum and energy,
333 systems which contain a metal nanoparticle embedded in a significant
334 volume of solvent will still experience nanoparticle diffusion inside
335 the solvent droplet. To aid in measuring an accurate temperature
336 profile for these systems, a single gold atom at the origin of the
337 coordinate system was assigned a mass $10,000 \times$ its original
338 mass. The bonded and nonbonded interactions for this atom remain
339 unchanged and the heavy atom is excluded from the RNEMD velocity
340 scaling. The only effect of this gold atom is to effectively pin the
341 nanoparticle at the origin of the coordinate system, thereby
342 preventing translational diffusion of the nanoparticle due to Brownian
343 motion.
344
345 To provide statistical independence, five separate configurations were
346 simulated for each particle radius and ligand length. The
347 structures were unique, starting at the point of ligand placement,
348 in order to sample multiple surface-ligand configurations.
349
350
351 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352 % EFFECT OF PARTICLE SIZE
353 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354 \section{Results}
355
356 We modeled four sizes of nanoparticles ($R =$ 10, 15, 20, and 25
357 \AA). The smallest particle size produces the lowest interfacial
358 thermal conductance values for most of the of protecting groups
359 (Fig. \ref{fig:NPthiols_G}). Between the other three sizes of
360 nanoparticles, there is no discernible dependence of the interfacial
361 thermal conductance on the nanoparticle size. It is likely that the
362 differences in local curvature of the nanoparticle sizes studied here
363 do not disrupt the ligand packing and behavior in drastically
364 different ways.
365
366 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
367 % EFFECT OF LIGAND CHAIN LENGTH
368 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
369
370 We have also utilized half-monolayers of three lengths of
371 alkanethiolate ligands -- S(CH$_2$)$_3$CH$_3$, S(CH$_2$)$_7$CH$_3$,
372 and S(CH$_2$)$_{11}$CH$_3$ -- referred to as C$_4$, C$_8$, and
373 C$_{12}$ respectively, in this study. Unlike our previous study of varying thiolate ligand chain lengths on
374 planar Au(111) surfaces, the interfacial thermal conductance of
375 ligand-protected nanospheres exhibits a distinct dependence on the
376 ligand length. For the three largest particle sizes, a half-monolayer
377 coverage of $C_4$ yields the highest interfacial thermal conductance
378 and the next-longest ligand, $C_8$, provides a similar boost. The
379 longest ligand, $C_{12}$, offers only a nominal ($\sim$ 10 \%)
380 increase in the interfacial thermal conductance over the bare
381 nanoparticles.
382
383 \begin{figure}
384 \includegraphics[width=\linewidth]{figures/NPthiols_G}
385 \caption{Interfacial thermal conductance ($G$) values for 4
386 sizes of solvated nanoparticles that are bare or protected
387 with a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$
388 alkanethiolate ligands.}
389 \label{fig:NPthiols_G}
390 \end{figure}
391
392 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
393 % HEAT TRANSFER MECHANISMS
394 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
395 \section{Mechanisms for Ligand-Enhanced Heat Transfer}
396
397 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398 % CORRUGATION OF PARTICLE SURFACE
399 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
400 \subsection{Corrugation of Particle Surface}
401
402 The bonding sites for thiols on gold surfaces have been studied
403 extensively and include configurations beyond the traditional atop,
404 bridge, and hollow sites found on planar surfaces. In particular, the
405 deep potential well between the gold atoms and the thiolate sulfurs
406 leads to insertion of the sulfur into the gold lattice and
407 displacement of interfacial gold atoms. The degree of ligand-induced
408 surface restructuring may have an impact on the interfacial thermal
409 conductance and is an important phenomenon to quantify.
410
411 Henz, \textit{et al.}\cite{Henz2007,Henz:2008qf} used the metal density as a
412 function of radius to measure the degree of mixing between the thiol
413 sulfurs and surface gold atoms at the edge of a nanoparticle. Although
414 metal density is important, disruption of the local crystalline
415 ordering would also have a large effect on the phonon spectrum in the
416 particles. To measure this effect, we use the fraction of gold atoms
417 exhibiting local fcc ordering as a function of radius to describe the
418 ligand-induced disruption of the nanoparticle surface.
419
420 The local bond orientational order can be described using the method
421 of Steinhardt \textit{et al.}\cite{Steinhardt1983} The local bonding environment, $\bar{q}_{\ell m}$, for each
422 atom in the system is determined by averaging over the spherical
423 harmonics between that atom and each of its neighbors,
424 \begin{equation}
425 \bar{q}_{\ell m} = \sum_i Y_\ell^m(\theta_i, \phi_i)
426 \end{equation}
427 where $\theta_i$ and $\phi_i$ are the relative angular coordinates of
428 neighbor $i$ in the laboratory frame. A global average orientational
429 bond order parameter, $\bar{Q}_{\ell m}$, is the average over each
430 $\bar{q}_{\ell m}$ for all atoms in the system. To remove the
431 dependence on the laboratory coordinate frame, the third order
432 rotationally invariant combination of $\bar{Q}_{\ell m}$,
433 $\hat{w}_\ell$, is utilized here.\cite{Steinhardt1983,Vardeman:2008fk}
434
435 For $\ell=4$, the ideal face-centered cubic (fcc), body-centered cubic
436 (bcc), hexagonally close-packed (hcp), and simple cubic (sc) local
437 structures exhibit $\hat{w}_4$ values of -0.159, 0.134, 0.159, and
438 0.159, respectively. Because $\hat{w}_4$ exhibits an extreme value for
439 fcc structures, it is ideal for measuring local fcc
440 ordering. The spatial distribution of $\hat{w}_4$ local bond
441 orientational order parameters, $p(\hat{w}_4 , r)$, can provide
442 information about the location of individual atoms that are central to
443 local fcc ordering.
444
445 The fraction of fcc-ordered gold atoms at a given radius in the
446 nanoparticle,
447 \begin{equation}
448 f_\mathrm{fcc}(r) = \int_{-\infty}^{w_c} p(\hat{w}_4, r) d \hat{w}_4
449 \end{equation}
450 is described by the distribution of the local bond orientational order
451 parameters, $p(\hat{w}_4, r)$, and $w_c$, a cutoff for the peak
452 $\hat{w}_4$ value displayed by fcc structures. A $w_c$ value of -0.12
453 was chosen to isolate the fcc peak in $\hat{w}_4$.
454
455 As illustrated in Figure \ref{fig:Corrugation}, the presence of
456 ligands decreases the fcc ordering of the gold atoms at the
457 nanoparticle surface. For the smaller nanoparticles, this disruption
458 extends into the core of the nanoparticle, indicating widespread
459 disruption of the lattice.
460
461 \begin{figure}
462 \includegraphics[width=\linewidth]{figures/NP10_fcc}
463 \caption{Fraction of gold atoms with fcc ordering as a
464 function of radius for a 10 \AA\ radius nanoparticle. The
465 decreased fraction of fcc-ordered atoms in ligand-protected
466 nanoparticles relative to bare particles indicates
467 restructuring of the nanoparticle surface by the thiolate
468 sulfur atoms.}
469 \label{fig:Corrugation}
470 \end{figure}
471
472 We may describe the thickness of the disrupted nanoparticle surface by
473 defining a corrugation factor, $c$, as the ratio of the radius at
474 which the fraction of gold atoms with fcc ordering is 0.9 and the
475 radius at which the fraction is 0.5.
476
477 \begin{equation}
478 c = 1 - \frac{r(f_\mathrm{fcc} = 0.9)}{r(f_\mathrm{fcc} = 0.5)}
479 \end{equation}
480
481 A sharp interface will have a steep drop in $f_\mathrm{fcc}$ at the
482 edge of the particle ($c \rightarrow$ 0). In the opposite limit where
483 the entire nanoparticle surface is restructured by ligands, the radius
484 at which there is a high probability of fcc ordering moves
485 dramatically inward ($c \rightarrow$ 1).
486
487 The computed corrugation factors are shown in Figure
488 \ref{fig:NPthiols_combo} for bare nanoparticles and for
489 ligand-protected particles as a function of ligand chain length. The
490 largest nanoparticles are only slightly restructured by the presence
491 of ligands on the surface, while the smallest particle ($r$ = 10 \AA)
492 exhibits significant disruption of the original fcc ordering when
493 covered with a half-monolayer of thiol ligands.
494
495 Because the thiolate ligands do not significantly alter the larger
496 particle crystallinity, the surface corrugation does not seem to be a
497 likely candidate to explain the large increase in thermal conductance
498 at the interface when ligands are added.
499
500 % \begin{equation}
501 % C = \frac{r_{bare}(\rho_{\scriptscriptstyle{0.85}}) - r_{capped}(\rho_{\scriptscriptstyle{0.85}})}{r_{bare}(\rho_{\scriptscriptstyle{0.85}})}.
502 % \end{equation}
503 %
504 % 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.
505
506
507
508 \begin{figure}
509 \includegraphics[width=\linewidth]{figures/NPthiols_combo}
510 \caption{Computed corrugation values, solvent escape rates,
511 ligand orientational $P_2$ values, and interfacial solvent
512 orientational $P_2$ values for 4 sizes of solvated
513 nanoparticles that are bare or protected with a 50\%
514 coverage of C$_{4}$, C$_{8}$, or C$_{12}$ alkanethiolate
515 ligands.}
516 \label{fig:NPthiols_combo}
517 \end{figure}
518
519
520 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521 % MOBILITY OF INTERFACIAL SOLVENT
522 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
523 \subsection{Mobility of Interfacial Solvent}
524
525 Another possible mechanism for increasing interfacial conductance is
526 the mobility of the interfacial solvent. We used a survival
527 correlation function, $C(t)$, to measure the residence time of a
528 solvent molecule in the nanoparticle thiolate
529 layer.\cite{Stocker:2013cl} This function correlates the identity of
530 all hexane molecules within the radial range of the thiolate layer at
531 two separate times. If the solvent molecule is present at both times,
532 the configuration contributes a $1$, while the absence of the molecule
533 at the later time indicates that the solvent molecule has migrated
534 into the bulk, and this configuration contributes a $0$. A steep decay
535 in $C(t)$ indicates a high turnover rate of solvent molecules from the
536 chain region to the bulk. We may define the escape rate for trapped
537 solvent molecules at the interface as
538 \begin{equation}
539 k_\mathrm{escape} = \left( \int_0^T C(t) dt \right)^{-1}
540 \label{eq:mobility}
541 \end{equation}
542 where T is the length of the simulation. This is a direct measure of
543 the rate at which solvent molecules initially entangled in the
544 thiolate layer can escape into the bulk. When $k_\mathrm{escape}
545 \rightarrow 0$, the solvent becomes permanently trapped in the
546 interfacial region.
547
548 The solvent escape rates for bare and ligand-protected nanoparticles
549 are shown in Figure \ref{fig:NPthiols_combo}. As the ligand chain
550 becomes longer and more flexible, interfacial solvent molecules become
551 trapped in the ligand layer and the solvent escape rate decreases.
552 This mechanism contributes a partial explanation as to why the longer
553 ligands have significantly lower thermal conductance.
554
555 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556 % ORIENTATION OF LIGAND CHAINS
557 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
558 \subsection{Orientation of Ligand Chains}
559
560 As the ligand chain length increases in length, it exhibits
561 significantly more conformational flexibility. Thus, different lengths
562 of ligands should favor different chain orientations on the surface of
563 the nanoparticle. To determine the distribution of ligand orientations
564 relative to the particle surface we examine the probability of
565 finding a ligand with a particular orientation relative to the surface
566 normal of the nanoparticle,
567 \begin{equation}
568 \cos{(\theta)}=\frac{\vec{r}_i\cdot\hat{u}_i}{|\vec{r}_i||\hat{u}_i|}
569 \end{equation}
570 where $\vec{r}_{i}$ is the vector between the cluster center of mass
571 and the sulfur atom on ligand molecule {\it i}, and $\hat{u}_{i}$ is
572 the vector between the sulfur atom and \ce{CH3} pseudo-atom on ligand
573 molecule {\it i}. As depicted in Figure \ref{fig:NP_pAngle}, $\theta
574 \rightarrow 180^{\circ}$ for a ligand chain standing upright on the
575 particle ($\cos{(\theta)} \rightarrow -1$) and $\theta \rightarrow
576 90^{\circ}$ for a ligand chain lying down on the surface
577 ($\cos{(\theta)} \rightarrow 0$). As the thiolate alkane chain
578 increases in length and becomes more flexible, the ligands are more
579 willing to lie down on the nanoparticle surface and exhibit increased
580 population at $\cos{(\theta)} = 0$.
581
582 \begin{figure}
583 \includegraphics[width=\linewidth]{figures/NP_pAngle}
584 \caption{The two extreme cases of ligand orientation relative
585 to the nanoparticle surface: the ligand completely
586 outstretched ($\cos{(\theta)} = -1$) and the ligand fully
587 lying down on the particle surface ($\cos{(\theta)} = 0$).}
588 \label{fig:NP_pAngle}
589 \end{figure}
590
591 % \begin{figure}
592 % \includegraphics[width=\linewidth]{figures/thiol_pAngle}
593 % \caption{}
594 % \label{fig:thiol_pAngle}
595 % \end{figure}
596
597 An order parameter describing the average ligand chain orientation relative to
598 the nanoparticle surface is available using the second order Legendre
599 parameter,
600 \begin{equation}
601 P_2 = \left< \frac{1}{2} \left(3\cos^2(\theta) - 1 \right) \right>
602 \end{equation}
603
604 Ligand populations that are perpendicular to the particle surface have
605 $P_2$ values of 1, while ligand populations lying flat on the
606 nanoparticle surface have $P_2$ values of $-0.5$. Disordered ligand
607 layers will exhibit mean $P_2$ values of 0. As shown in Figure
608 \ref{fig:NPthiols_combo} the ligand $P_2$ values approaches 0 as
609 ligand chain length -- and ligand flexibility -- increases.
610
611 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612 % ORIENTATION OF INTERFACIAL SOLVENT
613 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
614 \subsection{Orientation of Interfacial Solvent}
615
616 Similarly, we examined the distribution of \emph{hexane} molecule
617 orientations relative to the particle surface using the same angular
618 analysis utilized for the ligand chain orientations. In this case,
619 $\vec{r}_i$ is the vector between the particle center of mass and one
620 of the \ce{CH2} pseudo-atoms in the middle of hexane molecule $i$ and
621 $\hat{u}_i$ is the vector between the two \ce{CH3} pseudo-atoms on
622 molecule $i$. Since we are only interested in the orientation of
623 solvent molecules near the ligand layer, we select only the hexane
624 molecules within a specific $r$-range, between the edge of the
625 particle and the end of the ligand chains. A large population of
626 hexane molecules with $\cos{(\theta)} \sim \pm 1$ would indicate
627 interdigitation of the solvent molecules between the upright ligand
628 chains. A more random distribution of $\cos{(\theta)}$ values
629 indicates a disordered arrangement of solvent molecules near the particle
630 surface. Again, $P_2$ order parameter values provide a population
631 analysis for the solvent that is close to the particle surface.
632
633 The average orientation of the interfacial solvent molecules is
634 notably flat on the \emph{bare} nanoparticle surfaces. This blanket of
635 hexane molecules on the particle surface may act as an insulating
636 layer, increasing the interfacial thermal resistance. As the length
637 (and flexibility) of the ligand increases, the average interfacial
638 solvent P$_2$ value approaches 0, indicating a more random orientation
639 of the ligand chains. The average orientation of solvent within the
640 $C_8$ and $C_{12}$ ligand layers is essentially random. Solvent
641 molecules in the interfacial region of $C_4$ ligand-protected
642 nanoparticles do not lie as flat on the surface as in the case of the
643 bare particles, but are not as randomly oriented as the longer ligand
644 lengths.
645
646 These results are particularly interesting in light of our previous
647 results\cite{Stocker:2013cl}, where solvent molecules readily filled
648 the vertical gaps between neighboring ligand chains and there was a
649 strong correlation between ligand and solvent molecular
650 orientations. It appears that the introduction of surface curvature
651 and a lower ligand packing density creates a disordered ligand layer
652 that lacks well-formed channels for the solvent molecules to occupy.
653
654 % \begin{figure}
655 % \includegraphics[width=\linewidth]{figures/hex_pAngle}
656 % \caption{}
657 % \label{fig:hex_pAngle}
658 % \end{figure}
659
660 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
661 % SOLVENT PENETRATION OF LIGAND LAYER
662 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
663 \subsection{Solvent Penetration of Ligand Layer}
664
665 We may also determine the extent of ligand -- solvent interaction by
666 calculating the hexane density as a function of radius. Figure
667 \ref{fig:hex_density} shows representative radial hexane density
668 profiles for a solvated 25 \AA\ radius nanoparticle with no ligands,
669 and 50\% coverage of C$_{4}$, C$_{8}$, and C$_{12}$ thiolates.
670
671 \begin{figure}
672 \includegraphics[width=\linewidth]{figures/hex_density}
673 \caption{Radial hexane density profiles for 25 \AA\ radius
674 nanoparticles with no ligands (circles), C$_{4}$ ligands
675 (squares), C$_{8}$ ligands (triangles), and C$_{12}$ ligands
676 (diamonds). As ligand chain length increases, the nearby
677 solvent is excluded from the ligand layer. Some solvent is
678 present inside the particle $r_{max}$ location due to
679 faceting of the nanoparticle surface.}
680 \label{fig:hex_density}
681 \end{figure}
682
683 The differences between the radii at which the hexane surrounding the
684 ligand-covered particles reaches bulk density correspond nearly
685 exactly to the differences between the lengths of the ligand
686 chains. Beyond the edge of the ligand layer, the solvent reaches its
687 bulk density within a few angstroms. The differing shapes of the
688 density curves indicate that the solvent is increasingly excluded from
689 the ligand layer as the chain length increases.
690
691 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692 % DISCUSSION
693 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
694 \section{Discussion}
695
696 The chemical bond between the metal and the ligand introduces
697 vibrational overlap that is not present between the bare metal surface
698 and solvent. Thus, regardless of ligand chain length, the presence of
699 a half-monolayer ligand coverage yields a higher interfacial thermal
700 conductance value than the bare nanoparticle. The dependence of the
701 interfacial thermal conductance on ligand chain length is primarily
702 explained by increased ligand flexibility and a corresponding decrease
703 in solvent mobility away from the particles. The shortest and least
704 flexible ligand ($C_4$), which exhibits the highest interfacial
705 thermal conductance value, has a smaller range of available angles relative to
706 the surface normal and is least likely to trap solvent molecules
707 within the ligand layer. The longer $C_8$ and $C_{12}$ ligands have
708 increasingly disordered orientations and correspondingly lower solvent
709 escape rates.
710
711 When the ligands are less tightly packed, the cooperative
712 orientational ordering between the ligand and solvent decreases
713 dramatically and the conductive heat transfer model plays a much
714 smaller role in determining the total interfacial thermal
715 conductance. Thus, heat transfer into the solvent relies primarily on
716 the convective model, where solvent molecules pick up thermal energy
717 from the ligands and diffuse into the bulk solvent. This mode of heat
718 transfer is hampered by a slow solvent escape rate, which is clearly
719 present in the randomly ordered long ligand layers.
720
721 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
722 % **ACKNOWLEDGMENTS**
723 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
724 \begin{acknowledgement}
725 Support for this project was provided by the National Science Foundation
726 under grant CHE-1362211. Computational time was provided by the
727 Center for Research Computing (CRC) at the University of Notre Dame.
728 \end{acknowledgement}
729
730
731 \newpage
732
733 \bibliography{NPthiols}
734
735 \end{document}

Properties

Name Value
svn:executable