ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/NPthiols/NPthiols.tex
Revision: 4201
Committed: Fri Aug 1 18:55:09 2014 UTC (9 years, 10 months ago) by kstocke1
Content type: application/x-tex
File size: 37624 byte(s)
Log Message:
Edits addressing reviewer comments. Updated plots with new data.

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

Properties

Name Value
svn:executable