ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4083
Committed: Sat Mar 15 23:00:32 2014 UTC (11 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 41737 byte(s)
Log Message:
Some minor edits

File Contents

# Content
1 \documentclass[journal = jctcce, 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{A method for creating thermal and angular momentum fluxes in
24 non-periodic simulations}
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\\ Department of Chemistry and Biochemistry\\ University of Notre Dame\\ Notre Dame, Indiana 46556}
30
31 \begin{document}
32
33 \begin{tocentry}
34 % \includegraphics[width=3.2cm]{figures/npVSS} \includegraphics[width=3.7cm]{figures/NP20}
35 \includegraphics[width=9cm]{figures/TOC}
36 \end{tocentry}
37
38 \newcolumntype{A}{p{1.5in}}
39 \newcolumntype{B}{p{0.75in}}
40
41 % \author{Kelsey M. Stocker and J. Daniel
42 % Gezelter\footnote{Corresponding author. \ Electronic mail:
43 % gezelter@nd.edu} \\
44 % 251 Nieuwland Science Hall, \\
45 % Department of Chemistry and Biochemistry,\\
46 % University of Notre Dame\\
47 % Notre Dame, Indiana 46556}
48
49 %\date{\today}
50
51 %\maketitle
52
53 %\begin{doublespace}
54
55 \begin{abstract}
56
57 We present a new reverse non-equilibrium molecular dynamics (RNEMD)
58 method that can be used with non-periodic simulation cells. This
59 method applies thermal and/or angular momentum fluxes between two
60 arbitrary regions of the simulation, and is capable of creating
61 stable temperature and angular velocity gradients while conserving
62 total energy and angular momentum. One particularly useful
63 application is the exchange of kinetic energy between two concentric
64 spherical regions, which can be used to generate thermal transport
65 between nanoparticles and the solvent that surrounds them. The
66 rotational couple to the solvent (a measure of interfacial friction)
67 is also available via this method. As tests of the new method, we
68 have computed the thermal conductivities of gold nanoparticles and
69 water clusters, the interfacial thermal conductivity ($G$) of a
70 solvated gold nanoparticle, and the interfacial friction of a
71 variety of solvated gold nanostructures.
72
73 \end{abstract}
74
75 \newpage
76
77 %\narrowtext
78
79 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
80 % **INTRODUCTION**
81 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
82 \section{Introduction}
83
84 Non-equilibrium molecular dynamics (NEMD) methods impose a temperature
85 or velocity {\it gradient} on a
86 system,\cite{Ashurst:1975eu,Evans:1982oq,Erpenbeck:1984qe,Evans:1986nx,Vogelsang:1988qv,Maginn:1993kl,Hess:2002nr,Schelling:2002dp,Berthier:2002ai,Evans:2002tg,Vasquez:2004ty,Backer:2005sf,Jiang:2008hc,Picalek:2009rz}
87 and use linear response theory to connect the resulting thermal or
88 momentum {\it flux} to transport coefficients of bulk materials,
89 \begin{equation}
90 j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}, \hspace{0.5in}
91 J_z = \lambda \frac{\partial T}{\partial z}.
92 \end{equation}
93 Here, $\frac{\partial T}{\partial z}$ and $\frac{\partial
94 v_x}{\partial z}$ are the imposed thermal and momentum gradients,
95 and as long as the imposed gradients are relatively small, the
96 corresponding fluxes, $J_z$ and $j_z(p_x)$, have a linear relationship
97 to the gradients. The coefficients that provide this relationship
98 correspond to physical properties of the bulk material, either the
99 shear viscosity $(\eta)$ or thermal conductivity $(\lambda)$. For
100 systems which include phase boundaries or interfaces, however, it is
101 often unclear what gradient (or discontinuity) should be imposed at
102 the boundary between materials.
103
104 In contrast, reverse Non-Equilibrium Molecular Dynamics (RNEMD)
105 methods impose an unphysical {\it flux} between different regions or
106 ``slabs'' of the simulation
107 box.\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Patel:2005zm,Shenogina:2009ix,Tenney:2010rp,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
108 The system responds by developing a temperature or velocity {\it
109 gradient} between the two regions. The gradients which develop in
110 response to the applied flux have the same linear response
111 relationships to the transport coefficient of interest. Since the
112 amount of the applied flux is known exactly, and measurement of a
113 gradient is generally less complicated, imposed-flux methods typically
114 take shorter simulation times to obtain converged results. At
115 interfaces, the observed gradients often exhibit near-discontinuities
116 at the boundaries between dissimilar materials. RNEMD methods do not
117 need many trajectories to provide information about transport
118 properties, and they have become widely used to compute thermal and
119 mechanical transport in both homogeneous liquids and
120 solids~\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Tenney:2010rp}
121 as well as heterogeneous
122 interfaces.\cite{Patel:2005zm,Shenogina:2009ix,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
123
124 The strengths of specific algorithms for imposing the flux between two
125 different slabs of the simulation cell has been the subject of some
126 renewed interest. The original RNEMD approach used kinetic energy or
127 momentum exchange between particles in the two slabs, either through
128 direct swapping of momentum vectors or via virtual elastic collisions
129 between atoms in the two regions. There have been recent
130 methodological advances which involve scaling all particle velocities
131 in both slabs.\cite{Kuang:2010if,Kuang:2012fe} Constraint equations
132 can be simultaneously imposed to require the simulation to conserve
133 both total energy and total linear momentum. The most recent and
134 simplest of the velocity scaling approaches allows for simultaneous
135 shearing (to provide viscosity estimates) as well as scaling (to
136 provide information about thermal conductivity).\cite{Kuang:2012fe}
137
138 To date, the primary method of studying heat transfer at {\it curved}
139 nanoscale interfaces has involved transient non-equilibrium molecular
140 dynamics and temperature relaxation.\cite{Lervik:2009fk} RNEMD methods
141 have only been used in periodic simulation cells where the exchange
142 regions are physically separated along one of the axes of the
143 simulation cell. This limits the applicability to infinite planar
144 interfaces which are perpendicular to the applied flux. In order to
145 model steady-state non-equilibrium distributions for curved surfaces
146 (e.g. hot nanoparticles in contact with colder solvent), or for
147 regions that are not planar slabs, the method requires some
148 generalization for non-parallel exchange regions. In the following
149 sections, we present a new velocity shearing and scaling (VSS) RNEMD
150 algorithm which has been explicitly designed for non-periodic
151 simulations, and use the method to compute some thermal transport and
152 solid-liquid friction at the surfaces of spherical and ellipsoidal
153 nanoparticles.
154
155 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156 % **METHODOLOGY**
157 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158 \section{Velocity shearing and scaling (VSS) for non-periodic systems}
159
160 The original periodic VSS-RNEMD approach uses a series of simultaneous
161 velocity shearing and scaling exchanges between the two
162 slabs.\cite{Kuang:2012fe} This method imposes energy and linear
163 momentum conservation constraints while simultaneously creating a
164 desired flux between the two slabs. These constraints ensure that all
165 configurations are sampled from the same microcanonical (NVE)
166 ensemble.
167
168 \begin{figure}
169 \includegraphics[width=\linewidth]{figures/npVSS2}
170 \caption{Schematics of periodic (left) and non-periodic (right)
171 Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
172 flux is applied from region A to region B. Thermal gradients are
173 depicted by a color gradient. Linear or angular velocity gradients
174 are shown as arrows.}
175 \label{fig:VSS}
176 \end{figure}
177
178 We have extended the VSS method for use in {\it non-periodic}
179 simulations, in which the ``slabs'' have been generalized to two
180 separated regions of space. These regions could be defined as
181 concentric spheres (as in Figure \ref{fig:VSS}), or one of the regions
182 can be defined in terms of a dynamically changing ``hull'' comprising
183 the surface atoms of the cluster. This latter definition is identical
184 to the hull used in the Langevin Hull algorithm.\cite{Vardeman2011}
185 For the non-periodic variant, the constraints fix both the total
186 energy and total {\it angular} momentum of the system while
187 simultaneously imposing a thermal and angular momentum flux between
188 the two regions.
189
190 After a time interval of $\Delta t$, the particle velocities
191 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
192 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
193 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$). The
194 scalars $a$ and $b$ collectively provide a thermal exchange between
195 the two regions. One of the values is larger than 1, and the other
196 smaller. To conserve total energy and angular momentum, the values of
197 these two scalars are coupled. The vectors ($\mathbf{c}_a$ and
198 $\mathbf{c}_b$) provide a relative rotational shear to the velocities
199 of the particles within the two regions, and these vectors must also
200 be coupled to constrain the total angular momentum.
201
202 Once the values of the scaling and shearing factors are known, the
203 velocity changes are applied,
204 \begin{displaymath}
205 \begin{array}{rclcl}
206 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
207 \underline{\mathrm{rotational~shearing}} \\ \\
208 \mathbf{v}_i $~~~$\leftarrow &
209 a \left(\mathbf{v}_i - \langle \omega_a
210 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
211 \mathbf{v}_j $~~~$\leftarrow &
212 b \left(\mathbf{v}_j - \langle \omega_b
213 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
214 \end{array}
215 \end{displaymath}
216 Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
217 velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
218 (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
219 to their velocities. The amount of shear is governed by the imposed angular momentum flux,
220 $\mathbf{j}_r(\mathbf{L})$,
221 \begin{eqnarray}
222 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
223 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
224 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
225 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
226 \end{eqnarray}
227 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia
228 tensor for each of the two shells.
229
230 To simultaneously impose a thermal flux ($J_r$) between the shells we
231 use energy conservation constraints,
232 \begin{eqnarray}
233 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
234 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
235 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
236 \cdot \mathbf{c}_a \label{eq:Kc}\\
237 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
238 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
239 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
240 \end{eqnarray}
241 Simultaneous solution of these quadratic formulae for the scaling
242 coefficients, $a$ and $b$, will ensure that the simulation samples
243 from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is
244 the instantaneous translational kinetic energy of each shell. At each
245 time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
246 $\mathbf{c}_b$, subject to the imposed angular momentum flux,
247 $j_r(\mathbf{L})$, and kinetic energy flux, $J_r$, values. The new
248 particle velocities are computed, and the simulation continues. System
249 configurations after the transformations have exactly the same energy
250 ({\it and} angular momentum) as before the moves.
251
252 As the simulation progresses, the velocity transformations can be
253 performed on a regular basis, and the system will develop a
254 temperature and/or angular velocity gradient in response to the
255 applied flux. By fitting the radial temperature gradient, it is
256 straightforward to obtain the bulk thermal conductivity,
257 \begin{equation}
258 J_r = -\lambda \left( \frac{\partial T}{\partial r}\right)
259 \end{equation}
260 from the radial kinetic energy flux $(J_r)$ that was applied during
261 the simulation. In practice, it is significantly easier to use the
262 integrated form of Fourier's law for spherical geometries. In
263 sections \ref{sec:thermal} -- \ref{sec:rotation} we outline ways of
264 obtaining interfacial transport coefficients from these RNEMD
265 simulations.
266
267 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268 % **COMPUTATIONAL DETAILS**
269 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
270 \section{Computational Details}
271
272 The new VSS-RNEMD methodology for non-periodic system geometries has
273 been implemented in our group molecular dynamics code,
274 OpenMD.\cite{Meineke:2005gd,openmd} We have tested the new method to
275 calculate the thermal conductance of a gold nanoparticle and SPC/E
276 water cluster, and compared the results with previous bulk RNEMD and
277 experimental values. We have also investigated the interfacial thermal
278 conductance and interfacial rotational friction for gold
279 nanostructures solvated in hexane as a function of nanoparticle size
280 and shape.
281
282 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
283 % FORCE FIELD PARAMETERS
284 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
285 \subsection{Force fields}
286
287 Gold -- gold interactions are described by the quantum Sutton-Chen
288 (QSC) model.\cite{PhysRevB.59.3527} The QSC parameters are tuned to
289 experimental properties such as density, cohesive energy, and elastic
290 moduli and include zero-point quantum corrections.
291
292 The SPC/E water model~\cite{Berendsen87} is particularly useful for
293 validation of conductivities and shear viscosities. This model has
294 been used to previously test other RNEMD and NEMD approaches, and
295 there are reported values for thermal conductivies and shear
296 viscosities at a wide range of thermodynamic conditions that are
297 available for direct comparison.\cite{Bedrov:2000,Kuang:2010if}
298
299 Hexane molecules are described by the TraPPE united atom
300 model,\cite{TraPPE-UA.alkanes} which provides good computational
301 efficiency and reasonable accuracy for bulk thermal conductivity
302 values. In this model, sites are located at the carbon centers for
303 alkyl groups. Bonding interactions, including bond stretches and bends
304 and torsions, were used for intra-molecular sites closer than 3
305 bonds. For non-bonded interactions, Lennard-Jones potentials were
306 used. We have previously utilized both united atom (UA) and all-atom
307 (AA) force fields for thermal
308 conductivity,\cite{Kuang:2011ef,Kuang:2012fe,Stocker:2013cl} and since
309 the united atom force fields cannot populate the high-frequency modes
310 that are present in AA force fields, they appear to work better for
311 modeling thermal conductance at metal/ligand interfaces.
312
313 Gold -- hexane nonbonded interactions are governed by pairwise
314 Lennard-Jones parameters derived from Vlugt \emph{et
315 al}.\cite{vlugt:cpc2007154} They fitted parameters for the
316 interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the
317 effective potential of Hautman and Klein for the Au(111)
318 surface.\cite{hautman:4994}
319
320 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
321 % NON-PERIODIC DYNAMICS
322 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
323 % \subsection{Dynamics for non-periodic systems}
324 %
325 % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
326 % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
327 % compressibilities, which are typically problematic for traditional affine transform methods. We have had
328 % success applying this method to several different systems including bare metal nanoparticles, liquid water
329 % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
330 % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
331 % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
332 % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
333 % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
334 % subject to standard Newtonian dynamics.
335
336 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
337 % SIMULATION PROTOCOL
338 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
339 \subsection{Simulation protocol}
340
341 In all cases, systems were equilibrated under non-periodic
342 isobaric-isothermal (NPT) conditions -- using the Langevin Hull
343 methodology\cite{Vardeman2011} -- before any non-equilibrium methods
344 were introduced. For heterogeneous systems, the gold nanoparticles and
345 ellipsoids were created from a bulk fcc lattice and were thermally
346 equilibrated before being solvated in hexane. Packmol\cite{packmol}
347 was used to solvate previously equilibrated gold nanostructures within
348 a spherical droplet of liquid hexane.
349
350 Once equilibrated, thermal or angular momentum fluxes were applied for
351 1 - 2 ns, until stable temperature or angular velocity gradients had
352 developed. Systems containing liquids were run under moderate pressure
353 (5 atm) and temperatures (230 K) to avoid the formation of a vapor
354 layer at the boundary of the cluster. Pressure was applied to the
355 system via the non-periodic Langevin Hull.\cite{Vardeman2011} However,
356 coupling to the external temperature and pressure bath was
357 removed to avoid interference with the imposed RNEMD flux.
358
359 Because the method conserves \emph{total} angular momentum, systems
360 which contain a metal nanoparticle embedded in a significant volume of
361 solvent will still experience nanoparticle diffusion inside the
362 solvent droplet. To aid in computing the interfacial transport properties in these
363 systems, a single gold atom at the origin of the coordinate system was
364 assigned a mass $10,000 \times$ its original mass. The bonded and
365 nonbonded interactions for this atom remain unchanged and the heavy
366 atom is excluded from the RNEMD exchanges. The only effect of this
367 gold atom is to effectively pin the nanoparticle at the origin of the
368 coordinate system, while still allowing for any desired rotation about a given axis. For rotation of
369 the gold ellipsoids we added two of these heavy atoms along the axis
370 of rotation, separated by an equal distance from the origin of the
371 coordinate system. These heavy atoms prevent off-axis tumbling of the
372 nanoparticle and allow for measurement of rotational friction relative
373 to a particular axis of the ellipsoid.
374
375 Angular velocity data was collected for the heterogeneous systems
376 after a brief period of imposed flux to initialize rotation of the
377 solvated nanostructure. Doing so ensures that we overcome the initial
378 static friction and calculate only the dynamic interfacial rotational
379 friction.
380
381 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
382 % THERMAL CONDUCTIVITIES
383 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
384 \subsection{Thermal conductivities}
385 \label{sec:thermal}
386
387 To compute the thermal conductivities of bulk materials, the
388 integrated form of Fourier's Law of heat conduction in radial
389 coordinates can be used to obtain an expression for the heat transfer
390 rate between concentric spherical shells:
391 \begin{equation}
392 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
393 \label{eq:Q}
394 \end{equation}
395 The heat transfer rate, $q_r$, is constant in spherical geometries,
396 while the heat flux, $J_r$ depends on the surface area of the two
397 shells. $\lambda$ is the thermal conductivity, and $T_{a,b}$ and
398 $r_{a,b}$ are the temperatures and radii of the two concentric RNEMD
399 regions, respectively.
400
401 A kinetic energy flux is created using VSS-RNEMD moves, and the
402 temperature in a set of 2~\AA\ thick radial shells is recorded. The
403 resulting temperature profiles are analyzed to yield information about
404 the interfacial thermal conductance. As the simulation progresses,
405 the VSS moves are performed on a regular basis, and the system
406 develops a thermal or velocity gradient in response to the applied
407 flux. Once a stable thermal gradient has been established between the
408 two regions, the thermal conductivity, $\lambda$, can be calculated
409 using a linear regression of the thermal gradient, $\langle
410 \frac{dT}{dr} \rangle$:
411
412 \begin{equation}
413 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
414 \label{eq:lambda}
415 \end{equation}
416
417 The rate of heat transfer, $q_r$, between the two RNEMD regions is
418 easily obtained from either the applied kinetic energy flux and the
419 area of the smaller of the two regions, or from the total amount of
420 transferred kinetic energy and the run time of the simulation.
421 \begin{equation}
422 q_r = J_r \cdot A = \frac{\Delta KE}{\Delta t}
423 \label{eq:heat}
424 \end{equation}
425
426 \subsubsection{Thermal conductivity of nanocrystalline gold}
427 Calculated values for the thermal conductivity of a 40 \AA\ radius
428 gold nanoparticle (15707 atoms) at a range of kinetic energy flux
429 values are shown in Table \ref{table:goldTC}. For these calculations,
430 the hot and cold slabs were excluded from the linear regression of the
431 thermal gradient.
432
433 \begin{table}
434 \caption{Calculated thermal gradients of a crystalline gold
435 nanoparticle of radius 40 \AA\ subject to a range of applied
436 kinetic energy fluxes. Calculations were performed at an average
437 temperature of 300 K. Gold-gold interactions are described by the
438 Quantum Sutton-Chen potential.}
439 \label{table:goldTC}
440 \begin{tabular}{cc}
441 \\ \hline \hline
442 {$J_r$} & {$\langle dT/dr \rangle$} \\
443 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K /
444 \AA)} \\ \hline \\
445 3.25$\times 10^{-6}$ & 0.114 \\
446 6.50$\times 10^{-6}$ & 0.232 \\
447 1.30$\times 10^{-5}$ & 0.449 \\
448 3.25$\times 10^{-5}$ & 1.180 \\
449 6.50$\times 10^{-5}$ & 2.339 \\ \hline \hline
450 \end{tabular}
451 \end{table}
452
453 The measured thermal gradients $\langle dT/dr \rangle$ are linearly
454 dependent on the applied kinetic energy flux $J_r$. The calculated
455 thermal conductivity value, $\lambda = 1.004 \pm
456 0.009$~W~/~m~$\cdot$~K compares well with previous {\it bulk} QSC
457 values of 1.08~--~1.26~W~/~m~$\cdot$~K\cite{Kuang:2010if}, though
458 still significantly lower than the experimental value of
459 320~W~/~m~$\cdot$~K, as the QSC force field neglects the significant
460 electronic contributions to thermal conductivity. We note that there
461 is only a minimal effect on the computed thermal conductivity of gold
462 when comparing periodic (bulk) simulations carried out at the same
463 densities as those used here.
464
465 \subsubsection{Thermal conductivity of an SPC/E water droplet}
466
467 Calculated values for the thermal conductivity of a cluster of 6912
468 SPC/E water molecules were computed with a range of applied kinetic
469 energy fluxes. As with the gold nanoparticle, the measured slopes were
470 linearly dependent on the applied kinetic energy flux $J_r$, and the
471 RNEMD regions were excluded from the $\langle dT/dr \rangle$ fit (see
472 Figure \ref{fig:lambda_G}).
473
474 The computed mean value of the thermal conductivity, $\lambda = 0.884
475 \pm 0.013$~W~/~m~$\cdot$~K, compares well with previous
476 non-equilibrium molecular dynamics simulations of bulk SPC/E that were
477 carried out in periodic geometries.\cite{Zhang2005,Romer2012} These
478 simualtions gave conductivity values from 0.81--0.87~W~/~m~$\cdot$~K,
479 and all of the simulated values are approximately 30-45\% higher than the
480 experimental thermal conductivity of water, which has been measured at
481 0.61~W~/~m~$\cdot$~K.\cite{WagnerKruse}
482
483 \begin{figure}
484 \includegraphics[width=\linewidth]{figures/lambda_G}
485 \caption{Upper panel: temperature profile of a typical SPC/E water
486 droplet. The RNEMD simulation transfers kinetic energy from region
487 A to region B, and the system responds by creating a temperature
488 gradient (circles). The fit to determine $\langle dT/dr \rangle$ is
489 carried out between the two RNEMD exchange regions. Lower panel:
490 temperature profile for a solvated Au nanoparticle ($r = 20$ \AA).
491 The temperature differences used to compute the total Kapitza
492 resistance (and interfacial thermal conductance) are indicated with
493 arrows.}
494 \label{fig:lambda_G}
495 \end{figure}
496
497
498 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499 % INTERFACIAL THERMAL CONDUCTANCE
500 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501 \subsection{Interfacial thermal conductance}
502 \label{sec:interfacial}
503
504 The interfacial thermal conductance, $G$, of a heterogeneous interface
505 located at $r_0$ can be understood as the change in thermal
506 conductivity in a direction normal $(\mathbf{\hat{n}})$ to the
507 interface,
508 \begin{equation}
509 G = \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{r_0}
510 \end{equation}
511 For heterogeneous systems such as the solvated nanoparticle shown in
512 Figure \ref{fig:NP20}, the interfacial thermal conductance at the
513 surface of the nanoparticle can be determined using a kinetic energy
514 flux applied using the RNEMD method developed above.
515
516 In spherical geometries, it is most convenient to begin by finding the
517 Kapitza or interfacial thermal resistance for a thin spherical shell,
518 \begin{equation}
519 R_K = \frac{1}{G} = \frac{\Delta
520 T}{J_r}
521 \end{equation}
522 where $\Delta T$ is the temperature drop from the interior to the
523 exterior of the shell. For two concentric shells, the kinetic energy
524 flux ($J_r$) is not the same (as the surface areas are not the same),
525 but the heat transfer rate, $q_r = J_r \cdot A$ is constant. The thermal
526 resistance of a shell with interior radius $r$ is most conveniently
527 written as
528 \begin{equation}
529 R_K = \frac{1}{q_r} \Delta T 4 \pi r^2.
530 \label{eq:RK}
531 \end{equation}
532
533 \begin{figure}
534 \includegraphics[width=\linewidth]{figures/NP20}
535 \caption{A gold nanoparticle with a radius of 20 \AA\ solvated in
536 TraPPE-UA hexane. A kinetic energy flux is applied between the
537 nanoparticle and an outer shell of solvent to obtain the interfacial
538 thermal conductance, $G$, and the interfacial rotational resistance
539 $\Xi^{rr}$ is determined using an angular momentum flux. }
540 \label{fig:NP20}
541 \end{figure}
542
543 To model the thermal conductance across a wide interface (or multiple
544 interfaces) it is useful to consider concentric shells as resistors
545 wired in series. The resistance of the shells is then additive, and
546 the interfacial thermal conductance is the inverse of the total
547 Kapitza resistance:
548 \begin{equation}
549 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
550 T_i\right) 4 \pi r_i^2
551 \end{equation}
552 This series can be extended for any number of adjacent shells,
553 allowing for the calculation of the interfacial thermal conductance
554 for interfaces of considerable thickness, such as self-assembled
555 ligand monolayers on a metal surface.
556
557 \subsubsection{Interfacial thermal conductance of solvated gold
558 nanoparticles}
559 Calculated interfacial thermal conductance ($G$) values for three
560 sizes of gold nanoparticles and a flat Au(111) surface solvated in
561 TraPPE-UA hexane~\cite{Stocker:2013cl} are shown in Table
562 \ref{table:G}.
563
564 \begin{table}
565 \caption{Calculated interfacial thermal conductance ($G$) values for
566 gold nanoparticles of varying radii solvated in TraPPE-UA
567 hexane. The nanoparticle $G$ values are compared to previous
568 simulation results for a Au(111) interface in TraPPE-UA hexane.}
569 \label{table:G}
570 \begin{tabular}{cc}
571 \hline \hline
572 {Nanoparticle Radius} & {$G$}\\
573 {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
574 20 & {47.1 $\pm$ 5.0} \\
575 30 & {45.4 $\pm$ 1.2} \\
576 40 & {46.5 $\pm$ 2.1} \\
577 \hline
578 Au(111) & {30.2} \\
579 \hline \hline
580 \end{tabular}
581 \end{table}
582
583 The introduction of surface curvature increases the interfacial
584 thermal conductance by a factor of approximately $1.5$ relative to the
585 flat interface, although there is no apparent size-dependence in the
586 $G$ values obtained from our calculations. This is quite different
587 behavior than the size-dependence observed by Lervik \textit{et
588 al.}\cite{Lervik:2009fk} in their NEMD simulations of decane
589 droplets in water. In the liquid / liquid case, the surface tension is
590 strongly dependent on the droplet size, and the interfacial
591 resistivity increases with surface tension. In our case, the surface
592 tension of the solid / liquid interface does not have a strong
593 dependence on the particle radius at these particle sizes.
594
595 Simulations of larger nanoparticles may yet demonstrate limiting $G$
596 values close to the flat Au(111) slab, although any spherical particle
597 will have a significant fraction of the surface atoms in non-111
598 facets. It is not clear at this point whether the increase in thermal
599 conductance for the spherical particles is due to the increased
600 population of undercoordinated surface atoms, or to increased solid
601 angle space available for phonon scattering into the solvent. These
602 two possibilities do open up some interesting avenues for
603 investigation.
604
605 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
606 % INTERFACIAL ROTATIONAL FRICTION
607 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
608 \subsection{Interfacial rotational friction}
609 \label{sec:rotation}
610 The interfacial rotational resistance tensor, $\Xi^{rr}$, can be
611 calculated for heterogeneous nanostructure / solvent systems by applying
612 an angular momentum flux between the solvated nanostructure and a
613 spherical shell of solvent at the outer edge of the cluster. An
614 angular velocity gradient develops in response to the applied flux,
615 causing the nanostructure and solvent shell to rotate in opposite
616 directions about a given axis.
617
618 \begin{figure}
619 \includegraphics[width=\linewidth]{figures/E25-75}
620 \caption{A gold prolate ellipsoid of length 65 \AA\ and width 25
621 \AA\ solvated by TraPPE-UA hexane. An angular momentum flux is
622 applied between the ellipsoid and an outer shell of solvent.}
623 \label{fig:E25-75}
624 \end{figure}
625
626 Analytical solutions for the diagonal elements of the rotational
627 resistance tensor for a solvated spherical body of radius $r$ under
628 ideal stick boundary conditions can be estimated using Stokes' law
629 \begin{equation}
630 \Xi^{rr}_{stick} = 8 \pi \eta r^3
631 \label{eq:Xisphere}
632 \end{equation}
633 where $\eta$ is the dynamic viscosity of the surrounding solvent.
634
635 For general ellipsoids with semiaxes $\alpha$, $\beta$, and $\gamma$,
636 Perrin's extension of Stokes' law provides exact solutions for
637 symmetric prolate $(\alpha \geq \beta = \gamma)$ and oblate $(\alpha <
638 \beta = \gamma)$ ellipsoids under ideal stick conditions. The Perrin
639 elliptic integral parameter,
640 \begin{equation}
641 S = \frac{2}{\sqrt{\alpha^2 - \beta^2}} \ln \left[ \frac{\alpha + \sqrt{\alpha^2 - \beta^2}}{\beta} \right].
642 \label{eq:S}
643 \end{equation}
644
645 For a prolate ellipsoidal nanoparticle (see Figure \ref{fig:E25-75}),
646 the rotational resistance tensor $\Xi^{rr}$ is a $3 \times 3$ diagonal
647 matrix with elements
648 \begin{align}
649 \Xi^{rr}_\alpha =& \frac{32 \pi}{3} \eta \frac{ \left(
650 \alpha^2 - \beta^2 \right) \beta^2}{2\alpha - \beta^2 S}
651 \\ \nonumber \\
652 \Xi^{rr}_{\beta,\gamma} =& \frac{32 \pi}{3} \eta \frac{ \left( \alpha^4 - \beta^4 \right)}{ \left( 2\alpha^2 - \beta^2 \right)S - 2\alpha},
653 \label{eq:Xirr}
654 \end{align}
655 corresponding to rotation about the long axis ($\alpha$), and each of
656 the equivalent short axes ($\beta$ and $\gamma$), respectively.
657
658 Previous VSS-RNEMD simulations of the interfacial friction of the
659 planar Au(111) / hexane interface have shown that the flat interface
660 exists within slip boundary conditions.\cite{Kuang:2012fe} Hu and
661 Zwanzig\cite{Zwanzig} investigated the rotational friction
662 coefficients for spheroids under slip boundary conditions and obtained
663 numerial results for a scaling factor to be applied to
664 $\Xi^{rr}_{\mathit{stick}}$ as a function of ratio of the shorter
665 semiaxes to the longer. Under slip conditions, rotation of any
666 sphere and rotation of a prolate ellipsoid about its long axis does
667 not displace any solvent, so the resulting $\Xi^{rr}_{\mathit{slip}}$
668 approaches $0$. For rotation of a prolate ellipsoid (with the aspect
669 ratio shown here) about its short axis, Hu and Zwanzig obtained
670 $\Xi^{rr}_{\mathit{slip}}$ values of $35.9\%$ of the analytical
671 $\Xi^{rr}_{\mathit{stick}}$ result. This reduced rotational
672 resistance accounts for the reduced interfacial friction under slip
673 boundary conditions.
674
675 The simulated rotational friction coefficient,
676 $\Xi^{rr}_{\mathit{sim}}$, at the interface can be extracted from
677 non-periodic VSS-RNEMD simulations quite easily using the total
678 applied torque ($\tau$) and the observed angular velocity of the solid
679 structure ($\omega_\mathrm{solid}$),
680 \begin{equation}
681 \Xi^{rr}_{\mathit{sim}} = \frac{\tau}{\omega_\mathrm{solid}}
682 \label{eq:Xieff}
683 \end{equation}
684
685 The total applied torque required to overcome the interfacial friction
686 and maintain constant angular rotation of the solid particle is
687 \begin{equation}
688 \tau = \frac{j_r(\mathbf{L}) \cdot A}{2}
689 \label{eq:tau}
690 \end{equation}
691 where $j_r(\mathbf{L})$ is the applied angular momentum flux, $A$ is
692 the surface area of the solid nanoparticle, and the factor of $2$ is
693 present because the torque is exerted on both the particle and the
694 counter-rotating solvent shell.
695
696 \subsubsection{Rotational friction for gold nanostructures in hexane}
697
698 Table \ref{table:couple} shows the calculated rotational friction
699 coefficients $\Xi^{rr}$ for spherical gold nanoparticles and a prolate
700 ellipsoidal gold nanorod solvated in TraPPE-UA hexane. An angular
701 momentum flux was applied between an outer shell of solvent that was
702 at least 20 \AA\ away from the surface of the particle (region $A$)
703 and the gold nanostructure (region $B$). Dynamic viscosity estimates
704 $(\eta)$ for TraPPE-UA hexane under these particular temperature and
705 pressure conditions were determined by applying a traditional VSS-RNEMD
706 linear momentum flux to a periodic box of this solvent.
707
708 \begin{table}
709 \caption{Comparison of rotational friction coefficients under ideal stick
710 ($\Xi^{rr}_{\mathit{stick}}$) and
711 slip ($\Xi^{rr}_{\mathit{slip}}$) conditions, along with the simulated
712 ($\Xi^{rr}_{\mathit{sim}}$) rotational friction coefficients of gold
713 nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid
714 is oriented with the long axis along the $z$ direction.}
715 \label{table:couple}
716 \begin{tabular}{lccccc}
717 \\ \hline \hline
718 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{sim}}$} & {$\Xi^{rr}_{\mathit{sim}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
719 {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
720 Sphere (r = 20 \AA) & {$x = y = z$} & {3314} & 0 & {2386 $\pm$ 14} & {0.720 $\pm$ 0.004}\\
721 Sphere (r = 30 \AA) & {$x = y = z$} & {11749}& 0 & {8415 $\pm$ 274} & {0.716 $\pm$ 0.023}\\
722 Sphere (r = 40 \AA) & {$x = y = z$} & {34464}& 0 & {47544 $\pm$ 3051} & {1.380 $\pm$ 0.089}\\
723 Prolate Ellipsoid & {$x = y$} & {4991} & {1792} & {3128 $\pm$ 166} & {0.627 $\pm$ 0.033}\\
724 Prolate Ellipsoid & {$z$} & {1993} & 0 & {1590 $\pm$ 30} & {0.798 $\pm$ 0.015}
725 \\ \hline \hline
726 \end{tabular}
727 \end{table}
728
729 The results for $\Xi^{rr}_{\mathit{sim}}$ show that, in contrast with
730 the flat Au(111) / hexane interface, gold nanostructures solvated by
731 hexane are closer to stick than slip boundary conditions. At these
732 length scales, the nanostructures are not perfect spheroids due to
733 atomic `roughening' of the surface and therefore experience increased
734 interfacial friction, deviating significantly from the ideal slip
735 case. The 20 and 30~\AA\ radius nanoparticles experience approximately
736 70\% of the ideal stick boundary interfacial friction. Rotation of the
737 ellipsoid about its long axis more closely approaches the stick limit
738 than rotation about the short axis, which at first seems
739 counterintuitive. However, the `propellor' motion caused by rotation
740 around the short axis may exclude solvent from the rotation cavity or
741 move a sufficient amount of solvent along with the gold that a smaller
742 interfacial friction is actually experienced.
743
744 The largest nanoparticle (40 \AA\ radius) appears to experience
745 interfacial friction in excess of ideal stick conditions. Although
746 the solvent velocity field does not appear in the calculation of
747 $\Xi^{rr}$, we do note that the smaller particles exhibit solvent
748 velocity fields that can be fit with the $v(r,\theta) = \sin\theta
749 \left( A r + B/r^2 \right)$ form first discovered by
750 Jeffery,\cite{Jeffery:1915fk} while the solvent velocity fields for
751 the $r=40$ \AA\ particle approaches the infinite fluid solution and
752 can be fit well with only the second term. The larger particle (and
753 the resulting larger solvent volume) appears to be transitioning to a
754 regime that approaches the infinite solvent solutions.
755
756 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
757 % **DISCUSSION**
758 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
759 \section{Discussion}
760
761 We have adapted the VSS-RNEMD methodology so that it can be used to
762 apply kinetic energy and angular momentum fluxes in explicitly
763 non-periodic simulation geometries. Non-periodic VSS-RNEMD preserves
764 the strengths of the original periodic variant, specifically Boltzmann
765 velocity distributions and minimal thermal anisotropy, while extending
766 the constraints to conserve total energy and total \emph{angular}
767 momentum. This method also maintains the ability to impose the kinetic
768 energy and angular momentum fluxes either jointly or individually.
769
770 This method enables steady-state calculation of interfacial thermal
771 conductance and interfacial rotational friction in heterogeneous
772 clusters. We have demonstrated the abilities of the method on some
773 familiar systems (nanocrystalline gold, SPC/E water) as well as on
774 interfaces that have been studied only in planar geometries. For the
775 bulk systems that are well-understood in periodic geometries, the
776 non-periodic VSS-RNEMD provides very similar estimates of the thermal
777 conductivity to other simulation methods.
778
779 For nanoparticle-to-solvent heat transfer, we have observed that the
780 interfacial conductance exhibits a $1.5\times$ increase over the
781 planar Au(111) interface, although we do not see any dependence of the
782 conductance on the particle size. Because the surface tension effects
783 present in liquid/liquid heat transfer are not strong contributors
784 here, we suggest two possible mechanisms for the $1.5\times$ increase
785 over the planar surface: (1) the nanoparticles have an increased
786 population of undercoordinated surface atoms that are more efficient
787 at transferring vibrational energy to the solvent, or (2) there is
788 increased solid angle space available for phonon scattering into the
789 solvent. The second of these explanations would have a significant
790 dependence on the radius of spherical particles, although crystalline
791 faceting of the particles may be enough to reduce this dependence on
792 particle radius. These two possibilities open up some interesting
793 avenues for further exploration.
794
795 One area that this method opens up that was not available for periodic
796 simulation cells is direct computation of the solid / liquid
797 rotational friction coefficients of nanostructures. The systems we
798 have investigated (gold nanospheres and prolate ellipsoids) have
799 analytic stick and numerical slip solutions provided via
800 hydrodynamics, and our molecular simulations indicate that the bare
801 metallic nanoparticles are closer to stick than they are to slip
802 conditions. This is markedly different behavior than we have seen for
803 the solid / liquid friction at planar Au(111) interfaces in previous
804 simulations, where the solvents experience a nearly-laminar flow over
805 the surface. Schmidt and Skinner previously computed the behavior of
806 spherical tag particles in molecular dynamics simulations, and showed
807 that {\it slip} boundary conditions for translational resistance may
808 be more appropriate for molecule-sized spheres embedded in a sea of
809 spherical solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx} The
810 size of the gold nanoparticles studied here and the structuring of the
811 particle surfaces appear to bring the behavior closer to stick
812 boundaries. Exactly where the slip-stick crossover takes place is
813 also an interesting avenue for future exploration.
814
815 The ability to interrogate non-periodic effects (e.g. surface
816 curvature, edge effects between facets, and the splay of ligand
817 surface groups) on interfacial transport means that this method can be
818 applied to systems of broad experimental and theoretical interest.
819
820 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
821 % **ACKNOWLEDGMENTS**
822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
823 \begin{acknowledgement}
824 The authors thank Dr. Shenyu Kuang for helpful discussions. Support
825 for this project was provided by the National Science Foundation
826 under grant CHE-0848243. Computational time was provided by the
827 Center for Research Computing (CRC) at the University of Notre Dame.
828 \end{acknowledgement}
829
830
831 \newpage
832
833 \bibliography{nonperiodicVSS}
834
835 %\end{doublespace}
836 \end{document}

Properties

Name Value
svn:executable