ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/NPthiols/NPthiols.tex
Revision: 4159
Committed: Wed May 28 18:55:50 2014 UTC (10 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 37375 byte(s)
Log Message:
Latest edits

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 ($C_4$)
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 hampered by a slow solvent escape rate, which is was
66 observed in the longer-chain ligands.
67 \end{abstract}
68
69 \newpage
70
71 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72 % INTRODUCTION
73 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74 \section{Introduction}
75
76 The thermal properties of various nanostructured interfaces have 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 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 non-equilibrium 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 increase in 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 have been in 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 very efficient at moving
104 thermal energy away from the surface.\cite{Stocker:2013cl} Trapped
105 solvent molecules that were orientationally ordered with nearby
106 ligands (but which were less able to diffuse into the bulk) were able
107 to double the thermal conductance of the interface. This indicates
108 that the ligand-to-solvent vibrational energy transfer is the 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 the
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 effects those gaps have on the thermal conductance are 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 is 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 more applications of
161 ligand-stabilized nanostructures have become apparent, the structure
162 and dynamics of ligands on metallic nanoparticles have been studied
163 extensively.\cite{Badia1996:2,Badia1996,Henz2007,Henz:2008qf,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 To model thiolated gold nanospheres in this work, 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}. The nanoparticle / ligand complexes were
300 thermally equilibrated before Packmol was 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 configurations 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.
374
375 Unlike our previous study of varying thiolate ligand chain lengths on
376 planar Au(111) surfaces, the interfacial thermal conductance of
377 ligand-protected nanospheres exhibits a distinct dependence on the
378 ligand length. For the three largest particle sizes, a half-monolayer
379 coverage of $C_4$ yields the highest interfacial thermal conductance
380 and the next-longest ligand, $C_8$, provides a similar boost. The
381 longest ligand, $C_{12}$, offers only a nominal ($\sim$ 10 \%)
382 increase in the interfacial thermal conductance over the bare
383 nanoparticles.
384
385 \begin{figure}
386 \includegraphics[width=\linewidth]{figures/NPthiols_G}
387 \caption{Interfacial thermal conductance ($G$) values for 4
388 sizes of solvated nanoparticles that are bare or protected
389 with a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$
390 alkanethiolate ligands.}
391 \label{fig:NPthiols_G}
392 \end{figure}
393
394 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
395 % HEAT TRANSFER MECHANISMS
396 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
397 \section{Mechanisms for Ligand-Enhanced Heat Transfer}
398
399 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
400 % CORRUGATION OF PARTICLE SURFACE
401 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
402 \subsection{Corrugation of Particle Surface}
403
404 The bonding sites for thiols on gold surfaces have been studied
405 extensively and include configurations beyond the traditional atop,
406 bridge, and hollow sites found on planar surfaces. In particular, the
407 deep potential well between the gold atoms and the thiolate sulfurs
408 leads to insertion of the sulfur into the gold lattice and
409 displacement of interfacial gold atoms. The degree of ligand-induced
410 surface restructuring may have an impact on the interfacial thermal
411 conductance and is an important phenomenon to quantify.
412
413 Henz, \textit{et al.}\cite{Henz2007,Henz:2008qf} used the metal density as a
414 function of radius to measure the degree of mixing between the thiol
415 sulfurs and surface gold atoms at the edge of a nanoparticle. Although
416 metal density is important, disruption of the local crystalline
417 ordering would also have a large effect on the phonon spectrum in the
418 particles. To measure this effect, we use the fraction of gold atoms
419 exhibiting local fcc ordering as a function of radius to describe the
420 ligand-induced disruption of the nanoparticle surface.
421
422 The local bond orientational order can be described using the method
423 of Steinhardt \textit{et al.},\cite{Steinhardt1983} where spherical
424 harmonics are associated with a central atom and its nearest
425 neighbors. The local bonding environment, $\bar{q}_{\ell m}$, for each
426 atom in the system can be determined by averaging over the spherical
427 harmonics between the central atom and each of its neighbors,
428 \begin{equation}
429 \bar{q}_{\ell m} = \sum_i Y_\ell^m(\theta_i, \phi_i)
430 \end{equation}
431 where $\theta_i$ and $\phi_i$ are the relative angular coordinates of
432 neighbor $i$ in the laboratory frame. A global average orientational
433 bond order parameter, $\bar{Q}_{\ell m}$, is the average over each
434 $\bar{q}_{\ell m}$ for all atoms in the system. To remove the
435 dependence on the laboratory coordinate frame, the third order
436 rotationally invariant combination of $\bar{Q}_{\ell m}$,
437 $\hat{w}_\ell$, is utilized here.\cite{Steinhardt1983,Vardeman:2008fk}
438
439 For $\ell=4$, the ideal face-centered cubic (fcc), body-centered cubic
440 (bcc), hexagonally close-packed (hcp), and simple cubic (sc) local
441 structures exhibit $\hat{w}_4$ values of -0.159, 0.134, 0.159, and
442 0.159, respectively. Because $\hat{w}_4$ exhibits an extreme value for
443 fcc structures, this makes it ideal for measuring local fcc
444 ordering. The spatial distribution of $\hat{w}_4$ local bond
445 orientational order parameters, $p(\hat{w}_4 , r)$, can provide
446 information about the location of individual atoms that are central to
447 local fcc ordering.
448
449 The fraction of fcc-ordered gold atoms at a given radius in the
450 nanoparticle,
451 \begin{equation}
452 f_\mathrm{fcc}(r) = \int_{-\infty}^{w_c} p(\hat{w}_4, r) d \hat{w}_4
453 \end{equation}
454 is described by the distribution of the local bond orientational order
455 parameters, $p(\hat{w}_4, r)$, and $w_c$, a cutoff for the peak
456 $\hat{w}_4$ value displayed by fcc structures. A $w_c$ value of -0.12
457 was chosen to isolate the fcc peak in $\hat{w}_4$.
458
459 As illustrated in Figure \ref{fig:Corrugation}, the presence of
460 ligands decreases the fcc ordering of the gold atoms at the
461 nanoparticle surface. For the smaller nanoparticles, this disruption
462 extends into the core of the nanoparticle, indicating widespread
463 disruption of the lattice.
464
465 \begin{figure}
466 \includegraphics[width=\linewidth]{figures/NP10_fcc}
467 \caption{Fraction of gold atoms with fcc ordering as a
468 function of radius for a 10 \AA\ radius nanoparticle. The
469 decreased fraction of fcc-ordered atoms in ligand-protected
470 nanoparticles relative to bare particles indicates
471 restructuring of the nanoparticle surface by the thiolate
472 sulfur atoms.}
473 \label{fig:Corrugation}
474 \end{figure}
475
476 We may describe the thickness of the disrupted nanoparticle surface by
477 defining a corrugation factor, $c$, as the ratio of the radius at
478 which the fraction of gold atoms with fcc ordering is 0.9 and the
479 radius at which the fraction is 0.5.
480
481 \begin{equation}
482 c = 1 - \frac{r(f_\mathrm{fcc} = 0.9)}{r(f_\mathrm{fcc} = 0.5)}
483 \end{equation}
484
485 A sharp interface will have a sharp drop in $f_\mathrm{fcc}$ at the
486 edge of the particle ($c \rightarrow$ 0). In the opposite limit where
487 the entire nanoparticle surface is restructured by ligands, the radius
488 at which there is a high probability of fcc ordering moves
489 dramatically inward ($c \rightarrow$ 1).
490
491 The computed corrugation factors are shown in Figure
492 \ref{fig:NPthiols_combo} for bare nanoparticles and for
493 ligand-protected particles as a function of ligand chain length. The
494 largest nanoparticles are only slightly restructured by the presence
495 of ligands on the surface, while the smallest particle ($r$ = 10 \AA)
496 exhibits significant disruption of the original fcc ordering when
497 covered with a half-monolayer of thiol ligands.
498
499 Because the thiolate ligands do not significantly alter the larger
500 particle crystallinity, the surface corrugation does not seem to be a
501 likely candidate to explain the large increase in thermal conductance
502 at the interface.
503
504 % \begin{equation}
505 % C = \frac{r_{bare}(\rho_{\scriptscriptstyle{0.85}}) - r_{capped}(\rho_{\scriptscriptstyle{0.85}})}{r_{bare}(\rho_{\scriptscriptstyle{0.85}})}.
506 % \end{equation}
507 %
508 % 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.
509
510
511
512 \begin{figure}
513 \includegraphics[width=\linewidth]{figures/NPthiols_combo}
514 \caption{Computed corrugation values, solvent escape rates,
515 ligand orientational $P_2$ values, and interfacial solvent
516 orientational $P_2$ values for 4 sizes of solvated
517 nanoparticles that are bare or protected with a 50\%
518 coverage of C$_{4}$, C$_{8}$, or C$_{12}$ alkanethiolate
519 ligands.}
520 \label{fig:NPthiols_combo}
521 \end{figure}
522
523
524 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
525 % MOBILITY OF INTERFACIAL SOLVENT
526 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
527 \subsection{Mobility of Interfacial Solvent}
528
529 Another possible mechanism for increasing interfacial conductance is
530 the mobility of the interfacial solvent. We used a survival
531 correlation function, $C(t)$, to measure the residence time of a
532 solvent molecule in the nanoparticle thiolate
533 layer.\cite{Stocker:2013cl} This function correlates the identity of
534 all hexane molecules within the radial range of the thiolate layer at
535 two separate times. If the solvent molecule is present at both times,
536 the configuration contributes a $1$, while the absence of the molecule
537 at the later time indicates that the solvent molecule has migrated
538 into the bulk, and this configuration contributes a $0$. A steep decay
539 in $C(t)$ indicates a high turnover rate of solvent molecules from the
540 chain region to the bulk. We may define the escape rate for trapped
541 solvent molecules at the interface as
542 \begin{equation}
543 k_\mathrm{escape} = \left( \int_0^T C(t) dt \right)^{-1}
544 \label{eq:mobility}
545 \end{equation}
546 where T is the length of the simulation. This is a direct measure of
547 the rate at which solvent molecules initially entangled in the
548 thiolate layer can escape into the bulk. When $k_\mathrm{escape}
549 \rightarrow 0$, the solvent becomes permanently trapped in the
550 interfacial region.
551
552 The solvent escape rates for bare and ligand-protected nanoparticles
553 are shown in Figure \ref{fig:NPthiols_combo}. As the ligand chain
554 becomes longer and more flexible, interfacial solvent molecules become
555 trapped in the ligand layer and the solvent escape rate decreases.
556 This mechanism contributes a partial explanation as to why the longer
557 ligands have significantly lower thermal conductance.
558
559 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560 % ORIENTATION OF LIGAND CHAINS
561 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
562 \subsection{Orientation of Ligand Chains}
563
564 As the ligand chain length increases in length, it exhibits
565 significantly more conformational flexibility. Thus, different lengths
566 of ligands should favor different chain orientations on the surface of
567 the nanoparticle. To determine the distribution of ligand orientations
568 relative to the particle surface we examine the probability of
569 finding a ligand with a particular orientation relative to the surface
570 normal of the nanoparticle,
571 \begin{equation}
572 \cos{(\theta)}=\frac{\vec{r}_i\cdot\hat{u}_i}{|\vec{r}_i||\hat{u}_i|}
573 \end{equation}
574 where $\vec{r}_{i}$ is the vector between the cluster center of mass
575 and the sulfur atom on ligand molecule {\it i}, and $\hat{u}_{i}$ is
576 the vector between the sulfur atom and \ce{CH3} pseudo-atom on ligand
577 molecule {\it i}. As depicted in Figure \ref{fig:NP_pAngle}, $\theta
578 \rightarrow 180^{\circ}$ for a ligand chain standing upright on the
579 particle ($\cos{(\theta)} \rightarrow -1$) and $\theta \rightarrow
580 90^{\circ}$ for a ligand chain lying down on the surface
581 ($\cos{(\theta)} \rightarrow 0$). As the thiolate alkane chain
582 increases in length and becomes more flexible, the ligands are more
583 willing to lie down on the nanoparticle surface and exhibit increased
584 population at $\cos{(\theta)} = 0$.
585
586 \begin{figure}
587 \includegraphics[width=\linewidth]{figures/NP_pAngle}
588 \caption{The two extreme cases of ligand orientation relative
589 to the nanoparticle surface: the ligand completely
590 outstretched ($\cos{(\theta)} = -1$) and the ligand fully
591 lying down on the particle surface ($\cos{(\theta)} = 0$).}
592 \label{fig:NP_pAngle}
593 \end{figure}
594
595 % \begin{figure}
596 % \includegraphics[width=\linewidth]{figures/thiol_pAngle}
597 % \caption{}
598 % \label{fig:thiol_pAngle}
599 % \end{figure}
600
601 An order parameter the average ligand chain orientation relative to
602 the nanoparticle surface is available using the second order Legendre
603 parameter,
604 \begin{equation}
605 P_2 = \left< \frac{1}{2} \left(3\cos^2(\theta) - 1 \right) \right>
606 \end{equation}
607
608 Ligand populations that are perpendicular to the particle surface hav
609 P$_2$ values of 1, while ligand populations lying flat on the
610 nanoparticle surface have P$_2$ values of $-0.5$. Disordered ligand
611 layers will exhibit mean P$_2$ values of 0. As shown in Figure
612 \ref{fig:NPthiols_combo} the ligand P$_2$ values approaches 0 as
613 ligand chain length -- and ligand flexibility -- increases.
614
615 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
616 % ORIENTATION OF INTERFACIAL SOLVENT
617 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618 \subsection{Orientation of Interfacial Solvent}
619
620 Similarly, we examined the distribution of \emph{hexane} molecule
621 orientations relative to the particle surface using the same angular
622 analysis utilized for the ligand chain orientations. In this case,
623 $\vec{r}_i$ is the vector between the particle center of mass and one
624 of the \ce{CH2} pseudo-atoms in the middle of hexane molecule $i$ and
625 $\hat{u}_i$ is the vector between the two \ce{CH3} pseudo-atoms on
626 molecule $i$. Since we are only interested in the orientation of
627 solvent molecules near the ligand layer, we select only the hexane
628 molecules within a specific $r$-range, between the edge of the
629 particle and the end of the ligand chains. A large population of
630 hexane molecules with $\cos{(\theta)} \cong \pm 1$ would indicate
631 interdigitation of the solvent molecules between the upright ligand
632 chains. A more random distribution of $\cos{(\theta)}$ values
633 indicates a disordered arrangement of solvent chains on the particle
634 surface. Again, P$_2$ order parameter values provide a population
635 analysis for the solvent that is close to the particle surface.
636
637 The average orientation of the interfacial solvent molecules is
638 notably flat on the \emph{bare} nanoparticle surfaces. This blanket of
639 hexane molecules on the particle surface may act as an insulating
640 layer, increasing the interfacial thermal resistance. As the length
641 (and flexibility) of the ligand increases, the average interfacial
642 solvent P$_2$ value approaches 0, indicating a more random orientation
643 of the ligand chains. The average orientation of solvent within the
644 $C_8$ and $C_{12}$ ligand layers is essentially random. Solvent
645 molecules in the interfacial region of $C_4$ ligand-protected
646 nanoparticles do not lie as flat on the surface as in the case of the
647 bare particles, but are not as randomly oriented as the longer ligand
648 lengths.
649
650 These results are particularly interesting in light of our previous
651 results\cite{Stocker:2013cl}, where solvent molecules readily filled
652 the vertical gaps between neighboring ligand chains and there was a
653 strong correlation between ligand and solvent molecular
654 orientations. It appears that the introduction of surface curvature
655 and a lower ligand packing density creates a disordered ligand layer
656 that lacks well-formed channels for the solvent molecules to occupy.
657
658 % \begin{figure}
659 % \includegraphics[width=\linewidth]{figures/hex_pAngle}
660 % \caption{}
661 % \label{fig:hex_pAngle}
662 % \end{figure}
663
664 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
665 % SOLVENT PENETRATION OF LIGAND LAYER
666 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
667 \subsection{Solvent Penetration of Ligand Layer}
668
669 We may also determine the extent of ligand -- solvent interaction by
670 calculating the hexane density as a function of radius. Figure
671 \ref{fig:hex_density} shows representative radial hexane density
672 profiles for a solvated 25 \AA\ radius nanoparticle with no ligands,
673 and 50\% coverage of C$_{4}$, C$_{8}$, and C$_{12}$ thiolates.
674
675 \begin{figure}
676 \includegraphics[width=\linewidth]{figures/hex_density}
677 \caption{Radial hexane density profiles for 25 \AA\ radius
678 nanoparticles with no ligands (circles), C$_{4}$ ligands
679 (squares), C$_{8}$ ligands (triangles), and C$_{12}$ ligands
680 (diamonds). As ligand chain length increases, the nearby
681 solvent is excluded from the ligand layer. Some solvent is
682 present inside the particle $r_{max}$ location due to
683 faceting of the nanoparticle surface.}
684 \label{fig:hex_density}
685 \end{figure}
686
687 The differences between the radii at which the hexane surrounding the
688 ligand-covered particles reaches bulk density correspond nearly
689 exactly to the differences between the lengths of the ligand
690 chains. Beyond the edge of the ligand layer, the solvent reaches its
691 bulk density within a few angstroms. The differing shapes of the
692 density curves indicate that the solvent is increasingly excluded from
693 the ligand layer as the chain length increases.
694
695 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
696 % DISCUSSION
697 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
698 \section{Discussion}
699
700 The chemical bond between the metal and the ligand introduces
701 vibrational overlap that is not present between the bare metal surface
702 and solvent. Thus, regardless of ligand chain length, the presence of
703 a half-monolayer ligand coverage yields a higher interfacial thermal
704 conductance value than the bare nanoparticle. The dependence of the
705 interfacial thermal conductance on ligand chain length is primarily
706 explained by increased ligand flexibility and a corresponding decrease
707 in solvent mobility away from the particles. The shortest and least
708 flexible ligand ($C_4$), which exhibits the highest interfacial
709 thermal conductance value, has a smaller range of angles relative to
710 the surface normal and is least likely to trap solvent molecules
711 within the ligand layer. The longer $C_8$ and $C_{12}$ ligands have
712 increasingly disordered orientations and correspondingly lower solvent
713 escape rates.
714
715 When the ligands are less tightly packed, the cooperative
716 orientational ordering between the ligand and solvent decreases
717 dramatically and the conductive heat transfer model plays a much
718 smaller role in determining the total interfacial thermal
719 conductance. Thus, heat transfer into the solvent relies primarily on
720 the convective model, where solvent molecules pick up thermal energy
721 from the ligands and diffuse into the bulk solvent. This mode of heat
722 transfer is hampered by a slow solvent escape rate, which is clearly
723 present in the randomly ordered long ligand layers.
724
725 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
726 % **ACKNOWLEDGMENTS**
727 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
728 \begin{acknowledgement}
729 Support for this project was provided by the National Science Foundation
730 under grant CHE-1362211. Computational time was provided by the
731 Center for Research Computing (CRC) at the University of Notre Dame.
732 \end{acknowledgement}
733
734
735 \newpage
736
737 \bibliography{NPthiols}
738
739 \end{document}

Properties

Name Value
svn:executable