ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/NPthiols/NPthiols.tex
Revision: 4152
Committed: Fri May 23 16:21:47 2014 UTC (10 years, 1 month ago) by kstocke1
Content type: application/x-tex
File size: 33355 byte(s)
Log Message:

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{The Thermal Conductance of Alkanethiolate-Protected Gold
24 Nanospheres: Effects of Curvature and Chain Length}
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 % \includegraphics[width=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
49 \end{abstract}
50
51 \newpage
52
53 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54 % INTRODUCTION
55 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56 \section{Introduction}
57
58 The thermal properties of various nanostructured interfaces have been
59 the subject of intense experimental
60 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}
61 and the interfacial thermal conductance is the principal quantity of
62 interest for understanding interfacial heat
63 transport.\cite{cahill:793} Because nanoparticles have a significant
64 fraction of their atoms at the particle / solvent interface, the
65 chemical details of these interfaces govern the thermal transport
66 properties.
67
68 Previously, reverse non-equilibrium molecular dynamics (RNEMD) methods
69 have been applied to calculate the interfacial thermal conductance at
70 flat (111) metal / organic solvent interfaces that had been chemically
71 protected by mixed-chain alkanethiolate groups.\cite{kuang:AuThl}
72 These simulations suggested an explanation for the increase in thermal
73 conductivity at alkanethiol-capped metal surfaces compared with bare
74 metal interfaces. Specifically, the chemical bond between the metal
75 and the ligand introduces a vibrational overlap that is not present
76 without the protecting group, and the overlap between the vibrational
77 spectra (metal to ligand, ligand to solvent) provides a mechanism for
78 rapid thermal transport across the interface. The simulations also
79 suggest that this phenomenon is a non-monotonic function of the
80 fractional coverage of the surface, as moderate coverages allow
81 diffusive heat transport of solvent molecules that have been in close
82 contact with the ligands.
83
84 Additionally, simulations of {\it mixed-chain} alkylthiolate surfaces
85 showed that entrapped solvent can be very efficient at moving thermal
86 energy away from the surface.\cite{Stocker:2013cl} Trapped solvent that
87 is orientationally coupled to the ordered ligands (and is less able to
88 diffuse into the bulk) were able to double the thermal conductance of
89 the interface.
90
91 Recently, we extended RNEMD methods for use in non-periodic geometries
92 by creating scaling/shearing moves between concentric regions of the
93 simulation.\cite{Stocker:2014qq} The primary reason for developing a
94 non-periodic variant of RNEMD is to investigate the role that {\it
95 curved} nanoparticle surfaces play in heat and mass transport. On
96 planar surfaces, we discovered that orientational ordering of surface
97 protecting ligands had a large effect on the heat conduction from the
98 metal to the solvent. Smaller nanoparticles have high surface
99 curvature that creates gaps in well-ordered self-assembled monolayers,
100 and the effect those gaps will have on the thermal conductance are
101 unknown.
102
103
104
105 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106 % INTERFACIAL THERMAL CONDUCTANCE OF METALLIC NANOPARTICLES
107 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
108 %\section{Interfacial Thermal Conductance of Metallic Nanoparticles}
109
110 For a solvated nanoparticle, there is a critical value for the
111 interfacial thermal conductance,
112 \begin{equation}
113 G_c = \frac{3 C_s \Lambda_s}{R C_p}
114 \end{equation}
115 which depends on the solvent heat capacity, $C_s$, solvent thermal
116 conductivity, $\Lambda_s$, particle radius, $R$, and nanoparticle heat
117 capacity, $C_p$.\cite{Wilson:2002uq} In the limit of infinite
118 interfacial thermal conductance, $G >> G_c$, cooling of the
119 nanoparticle is limited by the solvent properties, $C_s$ and
120 $\Lambda_s$. In the opposite limit ($G << G_c$), the heat dissipation
121 is controlled by the thermal conductance of the particle / fluid
122 interface. It is this regime with which we are concerned, where
123 properties of the interface may be tuned to manipulate the rate of
124 cooling for a solvated nanoparticle. Based on estimates of $G$ from
125 previous simulations of gold nanoparticles solvated in hexane and
126 experimental results for solvated nanostructures, it appears that we
127 are in the $G << G_c$ regime for gold nanoparticles with radii smaller
128 than 40 nm when solvated in hexane. The particles included in this
129 study are more than an order of magnitude smaller than this critical
130 radius, so the heat dissipation should be controlled entirely by the
131 surface features of the particle / ligand / solvent interface.
132
133 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134 % STRUCTURE OF SELF-ASSEMBLED MONOLAYERS ON NANOPARTICLES
135 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
136 \section{Structure of Self-Assembled Monolayers on Nanoparticles}
137
138 Though the ligand packing on planar surfaces is characterized for many
139 different ligands and surface facets, it is not obvious \emph{a
140 priori} how the same ligands will behave on the highly curved
141 surfaces of nanoparticles. Thus, as more applications of
142 ligand-stabilized nanostructures have become apparent, the structure
143 and dynamics of ligands on metallic nanoparticles have been studied
144 extensively.\cite{Badia1996:2,Badia1996,Henz2007,Badia1997:2,Badia1997,Badia2000}
145 Badia, \textit{et al.} used transmission electron microscopy to
146 determine that alkanethiol ligands on gold nanoparticles pack
147 approximately 30\% more densely than on planar Au(111)
148 surfaces.\cite{Badia1996:2} Subsequent experiments demonstrated that
149 even at full coverages, surface curvature creates voids between linear
150 ligand chains that can be filled via interdigitation of ligands on
151 neighboring particles.\cite{Badia1996} The molecular dynamics
152 simulations of Henz, \textit{et al.} indicate that at low coverages,
153 the thiolate alkane chains will lie flat on the nanoparticle
154 surface\cite{Henz2007} Above 90\% coverage, the ligands stand upright
155 and recover the rigidity and tilt angle displayed on planar
156 facets. Their simulations also indicate a high degree of mixing
157 between the thiolate sulfur atoms and surface gold atoms at high
158 coverages.
159
160 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161 % NON-PERIODIC VSS-RNEMD METHODOLOGY
162 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
163 \section{Non-Periodic Velocity Shearing and Scaling RNEMD Methodology}
164
165 Non-periodic VSS-RNEMD\cite{Stocker:2014qq} periodically applies a
166 series of velocity scaling and shearing moves at regular intervals to
167 impose a flux between two concentric spherical regions. To impose a thermal flux between the shells while conserving total kinetic energy and angular momentum we solve for scaling coefficients $a$ and $b$
168
169 \begin{eqnarray}
170 a = \sqrt{1 - \frac{q_r \Delta t}{K_a - K_a^r}}\\ \nonumber\\
171 b = \sqrt{1 + \frac{q_r \Delta t}{K_b - K_b^r}}
172 \end{eqnarray}
173
174 at each time interval subject to the imposed thermal flux, $q_r$, the instantaneous translational kinetic energy of each shell, $K_{\{a,b\}}$, and the instatanteous rotational kinetic energy of each shell, $K_{\{a,b\}}^r$.
175
176 The scaling coefficients are determined and the velocity changes are applied
177 \begin{eqnarray}
178 \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~~\:\\
179 \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.
180 \end{eqnarray}
181
182 Here $\left < \omega_a \right > \times \mathbf{r}_i$ is the velocity for particle $i$ due to the overal angular velocity of the $a$ shell. In the absence of an angular momentum flux, the angular velocity $\left < \omega_a \right >$ of the shell is near 0 and the resultant particle velocity is a nearly linear scaling of the initial velocity by the coefficient $a$ or $b$.
183
184 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185 % CALCULATING TRANSPORT PROPERTIES
186 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187 \section{Calculating Transport Properties from Non-Periodic VSS-RNEMD}
188
189 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
190 % INTERFACIAL THERMAL CONDUCTANCE
191 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192 \subsection{Interfacial Thermal Conductance}
193
194 As described in our earlier work, the thermal conductance of each spherical shell may be defined as the
195 inverse Kapitza resistance of the shell. To describe the thermal conductance of an
196 interface of considerable thickness -- such as the ligand layers shown
197 here -- we can sum the individual thermal resistances of each concentric
198 spherical shell to arrive at the total thermal resistance, or the
199 inverse of the total interfacial thermal conductance. Unlike the periodic case, the intermediate temperature terms remain in the final sum, requiring the use of a series of individual resistance terms:
200
201 \begin{equation}
202 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
203 T_i\right) 4 \pi r_i^2.
204 \end{equation}
205
206 The longest ligand considered here is in excess of 15 \AA\ in length, and we use 10 concentric spherical shells to describe the total interfacial thermal conductance of the ligand layer.
207
208 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
209 % COMPUTATIONAL DETAILS
210 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
211 \section{Computational Details}
212
213 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
214 % FORCE FIELDS
215 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
216 \subsection{Force Fields}
217
218 Gold -- gold interactions are described by the quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} Hexane solvent is described by the TraPPE united atom model,\cite{TraPPE-UA.alkanes} where sites are located at the carbon centers for alkyl groups. The TraPPE-UA model for hexane provides both computational efficiency and reasonable accuracy for bulk thermal conductivity values. Bonding interactions were used for intra-molecular sites closer than 3 bonds. Effective Lennard-Jones potentials were used for non-bonded interactions.
219
220 To describe the interactions between metal (Au) and non-metal atoms, potential energy terms were adapted from an adsorption study of alkyl thiols on gold surfaces by Vlugt, \textit{et al.}\cite{vlugt:cpc2007154} They fit an effective pair-wise Lennard-Jones form of potential parameters for the interaction between Au and pseudo-atoms CH$_x$ and S based on a well-established and widely-used effective potential of Hautman and Klein for the Au(111) surface.\cite{hautman:4994}
221
222 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223 % SIMULATION PROTOCOL
224 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225 \subsection{Simulation Protocol}
226
227 The various sized gold nanoparticles were created from a bulk fcc lattice and were thermally equilibrated prior to the addition of ligands. A 50\% coverage of ligands (based on coverages reported by Badia, \textit{et al.}\cite{Badia1996:2}) were placed on the surface of the equilibrated nanoparticles using Packmol\cite{packmol}. The nanoparticle / ligand complexes were briefly thermally equilibrated before Packmol was used to solvate the structures within a spherical droplet of hexane. The thickness of the solvent layer was chosen to be at least 1.5$\times$ the radius of the nanoparticle / ligand structure. The fully solvated system was equilibrated for at least 1 nanoseconds using the Langevin Hull to apply 50 atm of pressure and a target temperature of 250 K.
228
229 Once equilibrated, thermal fluxes were applied for
230 1 nanosecond, until stable temperature gradients had
231 developed. Systems were run under moderate pressure
232 (50 atm) and average temperature (250K) to maintain a compact solvent cluster and avoid formation of a vapor phase near the heated metal surface. Pressure was applied to the
233 system via the non-periodic Langevin Hull.\cite{Vardeman2011} However,
234 thermal coupling to the external temperature and pressure bath was
235 removed to avoid interference with the imposed RNEMD flux.
236
237 Because the method conserves \emph{total} angular momentum and energy, systems
238 which contain a metal nanoparticle embedded in a significant volume of
239 solvent will still experience nanoparticle diffusion inside the
240 solvent droplet. To aid in measuring an accurate temperature profile for these
241 systems, a single gold atom at the origin of the coordinate system was
242 assigned a mass $10,000 \times$ its original mass. The bonded and
243 nonbonded interactions for this atom remain unchanged and the heavy
244 atom is excluded from the RNEMD velocity scaling. The only effect of this
245 gold atom is to effectively pin the nanoparticle at the origin of the
246 coordinate system, thereby preventing translational diffusion of the nanoparticle due to Brownian motion.
247
248 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
249 % INTERFACIAL THERMAL CONDUCTANCE
250 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
251 \section{Interfacial Thermal Conductance}
252
253 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254 % EFFECT OF PARTICLE SIZE
255 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
256 \subsection{Effect of Particle Size}
257
258 We have modeled four sizes of nanoparticles ($r =$ 10, 15, 20, and 25 \AA). The smallest particle size produces the lowest interfacial thermal conductance value regardless of protecting group. Between the other three sizes of nanoparticles, there is no discernible dependence of the interfacial thermal conductance on the nanoparticle size. It is likely that the differences in local curvature of the nanoparticle sizes studied here do not disrupt the ligand packing and behavior in drastically different ways.
259
260 \begin{figure}
261 \includegraphics[width=\linewidth]{figures/NP25_C12h1}
262 \caption{A 25 \AA\ radius gold nanoparticle protected with a half-monolayer of TraPPE-UA dodecanethiolate (C$_{12}$) ligands and solvated in TraPPE-UA hexane. The interfacial thermal conductance is computed by applying a kinetic energy flux between the nanoparticle and an outer shell of solvent.}
263 \label{fig:NP25_C12h1}
264 \end{figure}
265
266 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267 % EFFECT OF LIGAND CHAIN LENGTH
268 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
269 \subsection{Effect of Ligand Chain Length}
270
271 We have utilized a half-monolayer of three lengths of alkanethiolate ligands -- S(CH$_2$)$_3$CH$_3$, S(CH$_2$)$_7$CH$_3$, and S(CH$_2$)$_{11}$CH$_3$ -- referred to as C$_4$, C$_8$, and C$_{12}$ respectively, in this study.
272
273 Unlike our previous study of varying thiolate ligand chain lengths on Au(111) surfaces, the interfacial thermal conductance of ligand-protected nanoparticles exhibits a distinct dependence on the ligand length. For the three largest particle sizes, a half-monolayer coverage of $C_4$ yields the highest interfacial thermal conductance and the next-longest ligand $C_8$ provides a nearly equivalent boost. The longest ligand $C_{12}$ offers only a nominal ($\sim$ 10 \%) increase in the interfacial thermal conductance over a bare nanoparticle.
274
275 \begin{figure}
276 \includegraphics[width=\linewidth]{figures/NPthiols_Gcombo}
277 \caption{Interfacial thermal conductance ($G$) and corrugation values for 4 sizes of solvated nanoparticles that are bare or protected with a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$ alkanethiolate ligands.}
278 \label{fig:NPthiols_Gcombo}
279 \end{figure}
280
281 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
282 % HEAT TRANSFER MECHANISMS
283 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
284 \section{Mechanisms for Heat Transfer}
285
286 \begin{figure}
287 \includegraphics[width=\linewidth]{figures/NPthiols_combo}
288 \caption{Computed solvent escape rates, ligand orientational $P_2$ values, and interfacial solvent orientational $P_2$ values for 4 sizes of solvated nanoparticles that are bare or protected with a 50\% coverage of C$_{4}$, C$_{8}$, or C$_{12}$ alkanethiolate ligands.}
289 \label{fig:NPthiols_combo}
290 \end{figure}
291
292 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
293 % CORRUGATION OF PARTICLE SURFACE
294 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
295 \subsection{Corrugation of Particle Surface}
296
297 The bonding sites for thiols on gold surfaces have been studied extensively and include configurations beyond the traditional atop, bridge, and hollow sites found on planar surfaces. In particular, the deep potential well between the gold atoms and the thiolate sulfurs leads to insertion of the sulfur into the gold lattice and displacement of interfacial gold atoms. The degree of ligand-induced surface restructuring may have an impact on the interfacial thermal conductance and is an important phenomenon to quantify.
298
299 Henz, \textit{et al.}\cite{Henz2007} used the metal density as a function of radius to measure the degree of mixing between the thiol sulfurs and surface gold atoms at the edge of a nanoparticle. Although metal density is important, disruption of the local crystalline ordering would also have a large effect on the phonon spectrum in the particles. To measure this effect, we use the fraction of gold atoms exhibiting local fcc ordering as a function of radius to describe the ligand-induced disruption of the nanoparticle surface.
300
301 The local bond orientational order can be described using the model proposed by Steinhardt \textit{et al.},\cite{Steinhardt1983} where spherical harmonics are associated with a central atom and its nearest neighbors. The local bonding environment, $\bar{q}_{\ell m}$, for each atom in the system can be determined by averaging over the spherical harmonics between the central atom and each of its neighbors. A global average orientational bond order parameter, $\bar{Q}_{\ell m}$, is the average over each $\bar{q}_{\ell m}$ for all atoms in the system. The third-order rotationally invariant combination of $\bar{Q}_{\ell m}$, $\hat{W}_4$, is utilized here. Ideal face-centered cubic (fcc), body-centered cubic (bcc), hexagonally close-packed (hcp), and simple cubic (sc), have values in the $\ell$ = 4 $\hat{W}$ invariant of -0.159, 0.134, 0.159, and 0.159, respectively. $\hat{W}_4$ has an extreme value for fcc structures, making it ideal for measuring local fcc order. The distribution of $\hat{W}_4$ local bond orientational order parameters, $p(\hat{W}_4)$, can provide information about individual atoms that are central to local fcc ordering.
302
303 The fraction of fcc ordered gold atoms at a given radius
304
305 \begin{equation}
306 f_{fcc} = \int_{-\infty}^{w_i} p(\hat{W}_4) d \hat{W}_4
307 \end{equation}
308
309 is described by the distribution of the local bond orientational order parameter, $p(\hat{W}_4)$, and $w_i$, a cutoff for the peak $\hat{W}_4$ value displayed by fcc structures. A $w_i$ value of -0.12 was chosen to isolate the fcc peak in $\hat{W}_4$.
310
311 As illustrated in Figure \ref{fig:Corrugation}, the presence of ligands decreases the fcc ordering of the gold atoms at the nanoparticle surface. For the smaller nanoparticles, this disruption extends into the core of the nanoparticle, indicating widespread disruption of the lattice.
312
313 \begin{figure}
314 \includegraphics[width=\linewidth]{figures/NP10_fcc}
315 \caption{Fraction of gold atoms with fcc ordering as a function of radius for a 10 \AA\ radius nanoparticle. The decreased fraction of fcc ordered atoms in ligand-protected nanoparticles relative to bare particles indicates restructuring of the nanoparticle surface by the thiolate sulfur atoms.}
316 \label{fig:Corrugation}
317 \end{figure}
318
319 We may describe the thickness of the disrupted nanoparticle surface by defining a corrugation factor, $c$, as the ratio of the radius at which the fraction of gold atoms with fcc ordering is 0.9 and the radius at which the fraction is 0.5.
320
321 \begin{equation}
322 c = 1 - \frac{r(f_{fcc} = 0.9)}{r(f_{fcc} = 0.5)}
323 \end{equation}
324
325 A clean, unstructured interface will have a sharp drop in $f_{fcc}$ at the edge of the particle ($c \rightarrow$ 0). In the opposite limit where the entire nanoparticle surface is restructured, the radius at which there is a high probability of fcc ordering moves dramatically inward ($c \rightarrow$ 1).
326
327 The computed corrugation factors are shown in Figure \ref{fig:NPthiols_Gcombo} for bare nanoparticles and for ligand-protected particles as a function of ligand chain length. The largest nanoparticles are only slightly restructured by the presence of ligands on the surface, while the smallest particle ($r$ = 10 \AA) exhibits significant disruption of the original fcc ordering when covered with a half-monolayer of thiol ligands.
328
329 % \begin{equation}
330 % C = \frac{r_{bare}(\rho_{\scriptscriptstyle{0.85}}) - r_{capped}(\rho_{\scriptscriptstyle{0.85}})}{r_{bare}(\rho_{\scriptscriptstyle{0.85}})}.
331 % \end{equation}
332 %
333 % 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.
334
335 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336 % MOBILITY OF INTERFACIAL SOLVENT
337 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
338 \subsection{Mobility of Interfacial Solvent}
339
340 We use a survival correlation function, $C(t)$, to measure the
341 residence time of a solvent molecule in the nanoparticle thiolate
342 layer.\cite{Stocker:2013cl} This function correlates the identity of all
343 hexane molecules within the radial range of the thiolate layer at two
344 separate times. If the solvent molecule is present at both times, the
345 configuration contributes a $1$, while the absence of the molecule at
346 the later time indicates that the solvent molecule has migrated into
347 the bulk, and this configuration contributes a $0$. A steep decay in
348 $C(t)$ indicates a high turnover rate of solvent molecules from the
349 chain region to the bulk. We may define the escape rate for trapped
350 solvent molecules at the interface as
351
352 \begin{equation}
353 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
354 \label{eq:mobility}
355 \end{equation}
356
357 where T is the length of the simulation. This is a direct measure of the rate at which solvent molecules initially entangled in the thiolate layer can escape into the bulk. As $k_{escape} \rightarrow 0$, the solvent becomes permanently trapped in the interfacial region.
358
359 The solvent escape rates for bare and ligand-protected nanoparticles are shown in Figure \ref{fig:NPthiols_combo}. As the ligand chain becomes longer and more flexible, interfacial solvent molecules becomes trapped in the ligand layer and the solvent escape rate decreases.
360
361 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
362 % ORIENTATION OF LIGAND CHAINS
363 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364 \subsection{Orientation of Ligand Chains}
365
366 As the ligand chain length increases in length, it becomes significantly more flexible. Thus, different lengths of ligands should favor different chain orientations on the surface of the nanoparticle. To determine the distribution of ligand orientations relative to the particle surface we examine the probability of each $\cos{(\theta)}$,
367
368 \begin{equation}
369 \cos{(\theta)}=\frac{\vec{r}_i\cdot\hat{u}_i}{|\vec{r}_i||\hat{u}_i|}
370 \end{equation}
371
372 where $\vec{r}_{i}$ is the vector between the cluster center of mass and the sulfur atom on ligand molecule {\it i}, and $\hat{u}_{i}$ is the vector between the sulfur atom and \ce{CH3} pseudo-atom on ligand molecule {\it i}. As depicted in Figure \ref{fig:NP_pAngle}, $\theta \rightarrow 180^{\circ}$ for a ligand chain standing upright on the particle ($\cos{(\theta)} \rightarrow -1$) and $\theta \rightarrow 90^{\circ}$ for a ligand chain lying down on the surface ($\cos{(\theta)} \rightarrow 0$). As the thiolate alkane chain increases in length and becomes more flexible, the ligands are more likely to lie down on the nanoparticle surface and there will be increased population at $\cos{(\theta)} = 0$.
373
374 \begin{figure}
375 \includegraphics[width=\linewidth]{figures/NP_pAngle}
376 \caption{The two extreme cases of ligand orientation relative to the nanoparticle surface: the ligand completely outstretched ($\cos{(\theta)} = -1$) and the ligand fully lying down on the particle surface ($\cos{(\theta)} = 0$).}
377 \label{fig:NP_pAngle}
378 \end{figure}
379
380 % \begin{figure}
381 % \includegraphics[width=\linewidth]{figures/thiol_pAngle}
382 % \caption{}
383 % \label{fig:thiol_pAngle}
384 % \end{figure}
385
386 A single number describing the average ligand chain orientation relative to the nanoparticle surface may be achieved by calculating a P$_2$ order parameter from the distribution of $\cos(\theta)$ values.
387
388 \begin{equation}
389 P_2(\cos(\theta)) = \left < \frac{1}{2} \left (3\cos^2(\theta) - 1 \right ) \right >
390 \end{equation}
391
392 A ligand chain that is perpendicular to the particle surface has a P$_2$ value of 1, while a ligand chain lying flat on the nanoparticle surface has a P$_2$ value of $-0.5$. Disordered ligand layers will exhibit a mean P$_2$ value of 0. As shown in Figure \ref{fig:NPthiols_combo} the ligand P$_2$ value approaches 0 as ligand chain length -- and ligand flexibility -- increases.
393
394 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
395 % ORIENTATION OF INTERFACIAL SOLVENT
396 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
397 \subsection{Orientation of Interfacial Solvent}
398
399 Similarly, we examined the distribution of \emph{hexane} molecule orientations relative to the particle surface using the same $\cos{(\theta)}$ analysis utilized for the ligand chain orientations. In this case, $\vec{r}_i$ is the vector between the particle center of mass and one of the \ce{CH2} pseudo-atoms in the middle of hexane molecule $i$ and $\hat{u}_i$ is the vector between the two \ce{CH3} pseudo-atoms on molecule $i$. Since we are only interested in the orientation of solvent molecules near the ligand layer, we select only the hexane molecules within a specific $r$-range, between the edge of the particle and the end of the ligand chains. A large population of hexane molecules with $\cos{(\theta)} \cong -1$ would indicate interdigitation of the solvent molecules between the upright ligand chains. A more random distribution of $\cos{(\theta)}$ values indicates either little penetration of the ligand layer by the solvent, or a very disordered arrangement of ligand chains on the particle surface. Again, P$_2$ order parameter values may be obtained from the distribution of $\cos(\theta)$ values.
400
401 The average orientation of the interfacial solvent molecules is notably flat on the \emph{bare} nanoparticle surface. This blanket of hexane molecules on the particle surface may act as an insulating layer, increasing the interfacial thermal resistance. As the length (and flexibility) of the ligand increases, the average interfacial solvent P$_2$ value approaches 0, indicating random orientation of the ligand chains. The average orientation of solvent within the $C_8$ and $C_{12}$ ligand layers is essentially totally random. Solvent molecules in the interfacial region of $C_4$ ligand-protected nanoparticles do not lie as flat on the surface as in the case of the bare particles, but are not as randomly oriented as the longer ligand lengths.
402
403 These results are particularly interesting in light of our previous results\cite{Stocker:2013cl}, where solvent molecules readily filled the vertical gaps between neighboring ligand chains and there was a strong correlation between ligand and solvent molecular orientations. It appears that the introduction of surface curvature and a lower ligand packing density creates a very disordered ligand layer that lacks well-formed channels for the solvent molecules to occupy.
404
405 % \begin{figure}
406 % \includegraphics[width=\linewidth]{figures/hex_pAngle}
407 % \caption{}
408 % \label{fig:hex_pAngle}
409 % \end{figure}
410
411 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
412 % SOLVENT PENETRATION OF LIGAND LAYER
413 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
414 \subsection{Solvent Penetration of Ligand Layer}
415
416 We may also determine the extent of ligand -- solvent interaction by calculating the hexane density as a function of $r$. Figure \ref{fig:hex_density} shows representative radial hexane density profiles for a solvated 25 \AA\ radius nanoparticle with no ligands, and 50\% coverage of C$_{4}$, C$_{8}$, and C$_{12}$ thiolates.
417
418 \begin{figure}
419 \includegraphics[width=\linewidth]{figures/hex_density}
420 \caption{Radial hexane density profiles for 25 \AA\ radius nanoparticles with no ligands (circles), C$_{4}$ ligands (squares), C$_{8}$ ligands (triangles), and C$_{12}$ ligands (diamonds). As ligand chain length increases, the nearby solvent is excluded from the ligand layer. Some solvent is present inside the particle $r_{max}$ location due to faceting of the nanoparticle surface.}
421 \label{fig:hex_density}
422 \end{figure}
423
424 The differences between the radii at which the hexane surrounding the ligand-covered particles reaches bulk density correspond nearly exactly to the differences between the lengths of the ligand chains. Beyond the edge of the ligand layer, the solvent reaches its bulk density within a few angstroms. The differing shapes of the density curves indicate that the solvent is increasingly excluded from the ligand layer as the chain length increases.
425
426 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
427 % DISCUSSION
428 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
429 \section{Discussion}
430
431 The chemical bond between the metal and the ligand introduces vibrational overlap that is not present between the bare metal surface and solvent. Thus, regardless of ligand chain length, the presence of a half-monolayer ligand coverage yields a higher interfacial thermal conductance value than the bare nanoparticle. The dependence of the interfacial thermal conductance on ligand chain length is primarily explained by increased ligand flexibility. The shortest and least flexible ligand ($C_4$), which exhibits the highest interfacial thermal conductance value, is oriented more normal to the particle surface than the longer ligands and is least likely to trap solvent molecules within the ligand layer. The longer $C_8$ and $C_{12}$ ligands have increasingly disordered average orientations and correspondingly lower solvent escape rates.
432
433 When the ligands are less tightly packed, the cooperative orientational ordering between the ligand and solvent decreases dramatically and the conductive heat transfer model plays a much smaller role in determining the total interfacial thermal conductance. Thus, heat transfer into the solvent relies primarily on the convective model, where solvent molecules pick up thermal energy from the ligands and diffuse into the bulk solvent. This mode of heat transfer is hampered by a slow solvent escape rate, which is clearly present in the randomly ordered long ligand layers.
434
435 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
436 % **ACKNOWLEDGMENTS**
437 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
438 \begin{acknowledgement}
439 Support for this project was provided by the National Science Foundation
440 under grant CHE-1362211. Computational time was provided by the
441 Center for Research Computing (CRC) at the University of Notre Dame.
442 \end{acknowledgement}
443
444
445 \newpage
446
447 \bibliography{NPthiols}
448
449 \end{document}

Properties

Name Value
svn:executable