ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4077
Committed: Fri Mar 14 18:49:24 2014 UTC (10 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 41467 byte(s)
Log Message:
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=3.6cm]{figures/NP20} \includegraphics[width=3.6cm]{figures/E25-75}
35 \end{tocentry}
36
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 demonstrations and tests of
68 the new method, we have computed the thermal conductivities of gold
69 nanoparticles and water clusters, the shear viscosity of a water
70 cluster, the interfacial thermal conductivity ($G$) of a solvated
71 gold nanoparticle and the interfacial friction of a variety of
72 solvated gold nanostructures.
73
74 \end{abstract}
75
76 \newpage
77
78 %\narrowtext
79
80 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
81 % **INTRODUCTION**
82 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83 \section{Introduction}
84
85 Non-equilibrium molecular dynamics (NEMD) methods impose a temperature
86 or velocity {\it gradient} on a
87 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}
88 and use linear response theory to connect the resulting thermal or
89 momentum {\it flux} to transport coefficients of bulk materials,
90 \begin{equation}
91 j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}, \hspace{0.5in}
92 J_z = \lambda \frac{\partial T}{\partial z}.
93 \end{equation}
94 Here, $\frac{\partial T}{\partial z}$ and $\frac{\partial
95 v_x}{\partial z}$ are the imposed thermal and momentum gradients,
96 and as long as the imposed gradients are relatively small, the
97 corresponding fluxes, $J_z$ and $j_z(p_x)$, have a linear relationship
98 to the gradients. The coefficients that provide this relationship
99 correspond to physical properties of the bulk material, either the
100 shear viscosity $(\eta)$ or thermal conductivity $(\lambda)$. For
101 systems which include phase boundaries or interfaces, it is often
102 unclear what gradient (or discontinuity) should be imposed at the
103 boundary between materials.
104
105 In contrast, reverse Non-Equilibrium Molecular Dynamics (RNEMD)
106 methods impose an unphysical {\it flux} between different regions or
107 ``slabs'' of the simulation
108 box.\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Patel:2005zm,Shenogina:2009ix,Tenney:2010rp,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
109 The system responds by developing a temperature or velocity {\it
110 gradient} between the two regions. The gradients which develop in
111 response to the applied flux have the same linear response
112 relationships to the transport coefficient of interest. Since the
113 amount of the applied flux is known exactly, and measurement of a
114 gradient is generally less complicated, imposed-flux methods typically
115 take shorter simulation times to obtain converged results. At
116 interfaces, the observed gradients often exhibit near-discontinuities
117 at the boundaries between dissimilar materials. RNEMD methods do not
118 need many trajectories to provide information about transport
119 properties, and they have become widely used to compute thermal and
120 mechanical transport in both homogeneous liquids and
121 solids~\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Tenney:2010rp}
122 as well as heterogeneous
123 interfaces.\cite{Patel:2005zm,Shenogina:2009ix,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
124
125 The strengths of specific algorithms for imposing the flux between two
126 different slabs of the simulation cell has been the subject of some
127 renewed interest. The original RNEMD approach used kinetic energy or
128 momentum exchange between particles in the two slabs, either through
129 direct swapping of momentum vectors or via virtual elastic collisions
130 between atoms in the two regions. There have been recent
131 methodological advances which involve scaling all particle velocities
132 in both slabs.\cite{Kuang:2010if,Kuang:2012fe} Constraint equations
133 are simultaneously imposed to require the simulation to conserve both
134 total energy and total linear momentum. The most recent and simplest
135 of the velocity scaling approaches allows for simultaneous shearing
136 (to provide viscosity estimates) as well as scaling (to provide
137 information about thermal conductivity).\cite{Kuang:2012fe}
138
139 To date, the primary method of studying heat transfer at {\it curved}
140 nanoscale interfaes has involved transient non-equilibrium molecular
141 dynamics and temperature relaxation.\cite{Lervik:2009fk} RNEMD methods
142 have only been used in periodic simulation cells where the exchange
143 regions are physically separated along one of the axes of the
144 simulation cell. This limits the applicability to infinite planar
145 interfaces which are perpendicular to the applied flux. In order to
146 model steady-state non-equilibrium distributions for curved surfaces
147 (e.g. hot nanoparticles in contact with colder solvent), or for
148 regions that are not planar slabs, the method requires some
149 generalization for non-parallel exchange regions. In the following
150 sections, we present a new velocity shearing and scaling (VSS) RNEMD
151 algorithm which has been explicitly designed for non-periodic
152 simulations, and use the method to compute some thermal transport and
153 solid-liquid friction at the surfaces of spherical and ellipsoidal
154 nanoparticles.
155
156 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
157 % **METHODOLOGY**
158 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
159 \section{Velocity shearing and scaling (VSS) for non-periodic systems}
160
161 The original periodic VSS-RNEMD approach uses a series of simultaneous
162 velocity shearing and scaling exchanges between the two
163 slabs.\cite{Kuang:2012fe} This method imposes energy and linear
164 momentum conservation constraints while simultaneously creating a
165 desired flux between the two slabs. These constraints ensure that all
166 configurations are sampled from the same microcanonical (NVE)
167 ensemble.
168
169 \begin{figure}
170 \includegraphics[width=\linewidth]{figures/npVSS2}
171 \caption{Schematics of periodic (left) and non-periodic (right)
172 Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
173 flux is applied from region B to region A. Thermal gradients are
174 depicted by a color gradient. Linear or angular velocity gradients
175 are shown as arrows.}
176 \label{fig:VSS}
177 \end{figure}
178
179 We have extended the VSS method for use in {\it non-periodic}
180 simulations, in which the ``slabs'' have been generalized to two
181 separated regions of space. These regions could be defined as
182 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
183 can be defined in terms of a dynamically changing ``hull'' comprising
184 the surface atoms of the cluster. This latter definition is identical
185 to the hull used in the Langevin Hull algorithm.\cite{Vardeman2011}
186 For the non-periodic variant, the constraints fix both the total
187 energy and total {\it angular} momentum of the system while
188 simultaneously imposing a thermal and angular momentum flux between
189 the two regions.
190
191 After a time interval of $\Delta t$, the particle velocities
192 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
193 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
194 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$). The
195 scalars $a$ and $b$ collectively provide a thermal exchange between
196 the two regions. One of the values is larger than 1, and the other
197 smaller. To conserve total energy and angular momentum, the values of
198 these two scalars are coupled. The vectors ($\mathbf{c}_a$ and
199 $\mathbf{c}_b$) provide a relative rotational shear to the velocities
200 of the particles within the two regions, and these vectors must also
201 be coupled to constrain the total angular momentum.
202
203 Once the values of the scaling and shearing factors are known, the
204 velocity changes are applied,
205 \begin{displaymath}
206 \begin{array}{rclcl}
207 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
208 \underline{\mathrm{rotational~shearing}} \\ \\
209 \mathbf{v}_i $~~~$\leftarrow &
210 a \left(\mathbf{v}_i - \langle \omega_a
211 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
212 \mathbf{v}_j $~~~$\leftarrow &
213 b \left(\mathbf{v}_j - \langle \omega_b
214 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
215 \end{array}
216 \end{displaymath}
217 Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
218 velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
219 (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
220 to their velocities. The amount of shear is governed by the imposed angular momentum flux,
221 $\mathbf{j}_r(\mathbf{L})$,
222 \begin{eqnarray}
223 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
224 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
225 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
226 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
227 \end{eqnarray}
228 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia
229 tensor for each of the two shells.
230
231 To simultaneously impose a thermal flux ($J_r$) between the shells we
232 use energy conservation constraints,
233 \begin{eqnarray}
234 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
235 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
236 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
237 \cdot \mathbf{c}_a \label{eq:Kc}\\
238 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
239 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
240 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
241 \end{eqnarray}
242 Simultaneous solution of these quadratic formulae for the scaling coefficients, $a$ and $b$, will ensure that
243 the simulation samples from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is the instantaneous
244 translational kinetic energy of each shell. At each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
245 $\mathbf{c}_b$, subject to the imposed angular momentum flux, $j_r(\mathbf{L})$, and thermal flux, $J_r$,
246 values. The new particle velocities are computed, and the simulation continues. System configurations after the
247 transformations have exactly the same energy ({\it and} angular momentum) as before the moves.
248
249 As the simulation progresses, the velocity transformations can be
250 performed on a regular basis, and the system will develop a
251 temperature and/or angular velocity gradient in response to the
252 applied flux. By fitting the radial temperature gradient, it is
253 straightforward to obtain the bulk thermal conductivity,
254 \begin{equation}
255 J_r = -\lambda \left( \frac{\partial T}{\partial r}\right)
256 \end{equation}
257 from the radial kinetic energy flux $(J_r)$ that was applied during
258 the simulation. In practice, it is significantly easier to use the
259 integrated form of Fourier's law for spherical geometries. In
260 sections \ref{sec:thermal} -- \ref{sec:rotation} we outline ways of
261 obtaining interfacial transport coefficients from these RNEMD
262 simulations.
263
264 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
265 % **COMPUTATIONAL DETAILS**
266 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267 \section{Computational Details}
268
269 The new VSS-RNEMD methodology for non-periodic system geometries has
270 been implemented in our group molecular dynamics code,
271 OpenMD.\cite{Meineke:2005gd,openmd} We have tested the new method to
272 calculate the thermal conductance of a gold nanoparticle and SPC/E
273 water cluster, and compared the results with previous bulk RNEMD
274 values, as well as experiment. We have also investigated the
275 interfacial thermal conductance and interfacial rotational friction
276 for gold nanostructures solvated in hexane as a function of
277 nanoparticle size and shape.
278
279 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
280 % FORCE FIELD PARAMETERS
281 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
282 \subsection{Force field}
283
284 Gold -- gold interactions are described by the quantum Sutton-Chen
285 (QSC) model.\cite{PhysRevB.59.3527} The QSC parameters are tuned to
286 experimental properties such as density, cohesive energy, and elastic
287 moduli and include zero-point quantum corrections.
288
289 The SPC/E water model~\cite{Berendsen87} is particularly useful for
290 validation of conductivities and shear viscosities. This model has
291 been used to previously test other RNEMD and NEMD approaches, and
292 there are reported values for thermal conductivies and shear
293 viscosities at a wide range of thermodynamic conditions that are
294 available for direct comparison.\cite{Bedrov:2000,Kuang:2010if}
295
296 Hexane molecules are described by the TraPPE united atom
297 model,\cite{TraPPE-UA.alkanes} which provides good computational
298 efficiency and reasonable accuracy for bulk thermal conductivity
299 values. In this model, sites are located at the carbon centers for
300 alkyl groups. Bonding interactions, including bond stretches and bends
301 and torsions, were used for intra-molecular sites closer than 3
302 bonds. For non-bonded interactions, Lennard-Jones potentials were
303 used. We have previously utilized both united atom (UA) and all-atom
304 (AA) force fields for thermal
305 conductivity,\cite{Kuang:2011ef,Kuang:2012fe,Stocker:2013cl} and since
306 the united atom force fields cannot populate the high-frequency modes
307 that are present in AA force fields, they appear to work better for
308 modeling thermal conductance at metal/ligand interfaces.
309
310 Gold -- hexane nonbonded interactions are governed by pairwise
311 Lennard-Jones parameters derived from Vlugt \emph{et
312 al}.\cite{vlugt:cpc2007154} They fitted parameters for the
313 interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the
314 effective potential of Hautman and Klein for the Au(111)
315 surface.\cite{hautman:4994}
316
317 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
318 % NON-PERIODIC DYNAMICS
319 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
320 % \subsection{Dynamics for non-periodic systems}
321 %
322 % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
323 % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
324 % compressibilities, which are typically problematic for traditional affine transform methods. We have had
325 % success applying this method to several different systems including bare metal nanoparticles, liquid water
326 % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
327 % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
328 % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
329 % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
330 % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
331 % subject to standard Newtonian dynamics.
332
333 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
334 % SIMULATION PROTOCOL
335 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336 \subsection{Simulation protocol}
337
338 In all cases, systems were equilibrated under non-periodic
339 isobaric-isothermal (NPT) conditions -- using the Langevin Hull
340 methodology\cite{Vardeman2011} -- before any non-equilibrium methods
341 were introduced. For heterogeneous systems, the gold nanoparticles and
342 ellipsoids were created from a bulk fcc lattice and were thermally
343 equilibrated before being solvated in hexane. Packmol\cite{packmol}
344 was used to solvate previously equilibrated gold nanostructures within
345 a spherical droplet of hexane.
346
347 Once equilibrated, thermal or angular momentum fluxes were applied for
348 1 - 2 ns, until stable temperature or angular velocity gradients had
349 developed. Systems containing liquids were run under moderate pressure
350 (5 atm) and temperatures (230 K) to avoid the formation of a vapor
351 layer at the boundary of the cluster. Pressure was applied to the
352 system via the non-periodic Langevin Hull.\cite{Vardeman2011} However,
353 thermal coupling to the external temperature and pressure bath was
354 removed to avoid interference with the imposed RNEMD flux.
355
356 Because the method conserves \emph{total} angular momentum, systems
357 which contain a metal nanoparticle embedded in a significant volume of
358 solvent will still experience nanoparticle diffusion inside the
359 solvent droplet. To aid in computing the rotational friction in these
360 systems, a single gold atom at the origin of the coordinate system was
361 assigned a mass $10,000 \times$ its original mass. The bonded and
362 nonbonded interactions for this atom remain unchanged and the heavy
363 atom is excluded from the RNEMD exchanges. The only effect of this
364 gold atom is to effectively pin the nanoparticle at the origin of the
365 coordinate system, while still allowing for rotation. For rotation of
366 the gold ellipsoids we added two of these heavy atoms along the axis
367 of rotation, separated by an equal distance from the origin of the
368 coordinate system. These heavy atoms prevent off-axis tumbling of the
369 nanoparticle and allow for measurement of rotational friction relative
370 to a particular axis of the ellipsoid.
371
372 Angular velocity data was collected for the heterogeneous systems
373 after a brief period of imposed flux to initialize rotation of the
374 solvated nanostructure. Doing so ensures that we overcome the initial
375 static friction and calculate only the \emph{dynamic} interfacial
376 rotational friction.
377
378 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
379 % THERMAL CONDUCTIVITIES
380 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
381 \subsection{Thermal conductivities}
382 \label{sec:thermal}
383
384 To compute the thermal conductivities of bulk materials, the
385 integrated form of Fourier's Law of heat conduction in radial
386 coordinates can be used to obtain an expression for the heat transfer
387 rate between concentric spherical shells:
388 \begin{equation}
389 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
390 \label{eq:Q}
391 \end{equation}
392 The heat transfer rate, $q_r$, is constant in spherical geometries,
393 while the heat flux, $J_r$ depends on the surface area of the two
394 shells. $\lambda$ is the thermal conductivity, and $T_{a,b}$ and
395 $r_{a,b}$ are the temperatures and radii of the two concentric RNEMD
396 regions, respectively.
397
398 A kinetic energy flux is created using VSS-RNEMD moves, and the
399 temperature in each of the radial shells is recorded. The resulting
400 temperature profiles are analyzed to yield information about the
401 interfacial thermal conductance. As the simulation progresses, the
402 VSS moves are performed on a regular basis, and the system develops a
403 thermal or velocity gradient in response to the applied flux. Once a
404 stable thermal gradient has been established between the two regions,
405 the thermal conductivity, $\lambda$, can be calculated using a linear
406 regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
407
408 \begin{equation}
409 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
410 \label{eq:lambda}
411 \end{equation}
412
413 The rate of heat transfer, $q_r$, between the two RNEMD regions is
414 easily obtained from either the applied kinetic energy flux and the
415 area of the smaller of the two regions, or from the total amount of
416 transferred kinetic energy and the run time of the simulation.
417 \begin{equation}
418 q_r = J_r A = \frac{KE}{t}
419 \label{eq:heat}
420 \end{equation}
421
422 \subsubsection{Thermal conductivity of nanocrystalline gold}
423 Calculated values for the thermal conductivity of a 40 \AA$~$ radius
424 gold nanoparticle (15707 atoms) at a range of kinetic energy flux
425 values are shown in Table \ref{table:goldTC}. For these calculations,
426 the hot and cold slabs were excluded from the linear regression of the
427 thermal gradient.
428
429 \begin{table}
430 \caption{Calculated thermal gradients of a crystalline gold
431 nanoparticle of radius 40 \AA\ subject to a range of applied
432 kinetic energy fluxes. Calculations were performed at an average
433 temperature of 300 K. Gold-gold interactions are described by the
434 Quantum Sutton-Chen potential.}
435 \label{table:goldTC}
436 \begin{tabular}{cc}
437 \\ \hline \hline
438 {$J_r$} & {$\langle dT/dr \rangle$} \\
439 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K /
440 \AA)} \\ \hline \\
441 3.25$\times 10^{-6}$ & 0.114 \\
442 6.50$\times 10^{-6}$ & 0.232 \\
443 1.30$\times 10^{-5}$ & 0.449 \\
444 3.25$\times 10^{-5}$ & 1.180 \\
445 6.50$\times 10^{-5}$ & 2.339 \\ \hline \hline
446 \end{tabular}
447 \end{table}
448
449 The measured thermal gradients $\langle dT/dr \rangle$ are linearly
450 dependent on the applied kinetic energy flux $J_r$. The calculated
451 thermal conductivity value, $\lambda = 1.004 \pm
452 0.009$~W~/~m~$\cdot$~K compares well with previous {\it bulk} QSC
453 values of 1.08~--~1.26~W~/~m~$\cdot$~K\cite{Kuang:2010if}, though
454 still significantly lower than the experimental value of
455 320~W~/~m~$\cdot$~K, as the QSC force field neglects the significant
456 electronic contributions to thermal conductivity. We note that there
457 is only a minimal effect on the computed thermal conductivity of gold
458 when comparing periodic (bulk) simulations carried out at the same
459 densities as those used here.
460
461 \subsubsection{Thermal conductivity of a droplet of SPC/E water}
462
463 Calculated values for the thermal conductivity of a cluster of 6912
464 SPC/E water molecules were computed with a range of applied kinetic
465 energy fluxes. As with the gold nanoparticle, the measured slopes were
466 linearly dependent on the applied kinetic energy flux $J_r$, and the
467 RNEMD regions were excluded from the $\langle dT/dr \rangle$ fit (see
468 Fig. \ref{fig:lambda_G}).
469
470 The computed mean value of the thermal conductivity, $\lambda = 0.884
471 \pm 0.013$~W~/~m~$\cdot$~K, compares well with previous
472 non-equilibrium molecular dynamics simulations of bulk SPC/E that were
473 carried out in periodic geometries.\cite{Zhang2005,Romer2012} These
474 simualtions gave conductivity values from 0.81--0.87~W~/~m~$\cdot$~K,
475 and all of the simulated values are approximately 30-45\% higher than the
476 experimental thermal conductivity of water, which has been measured at
477 0.61~W~/~m~$\cdot$~K.\cite{WagnerKruse}
478
479 \begin{figure}
480 \includegraphics[width=\linewidth]{figures/lambda_G}
481 \caption{Upper panel: temperature profile of a typical SPC/E water
482 droplet. The RNEMD simulation transfers kinetic energy from region
483 A to region B, and the system responds by creating a temperature
484 gradient (circles). The fit to determine $\langle dT/dr \rangle$ is
485 carried out between the two RNEMD exchange regions. Lower panel:
486 temperature profile for a solvated Au nanoparticle ($r = 20$\AA).
487 The temperature differences used to compute the total Kapitza
488 resistance (and interfacial thermal conductance) are indicated with
489 arrows.}
490 \label{fig:lambda_G}
491 \end{figure}
492
493
494 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
495 % INTERFACIAL THERMAL CONDUCTANCE
496 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
497 \subsection{Interfacial thermal conductance}
498 \label{sec:interfacial}
499
500 The interfacial thermal conductance, $G$, of a heterogeneous interface
501 located at $r_0$ can be understood as the change in thermal
502 conductivity in a direction normal $(\mathbf{\hat{n}})$ to the
503 interface,
504 \begin{equation}
505 G = \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{r_0}
506 \end{equation}
507 For heterogeneous systems such as the solvated nanoparticle shown in
508 Figure \ref{fig:NP20}, the interfacial thermal conductance at the
509 surface of the nanoparticle can be determined using a kinetic energy
510 flux applied using the RNEMD method developed above.
511
512 In spherical geometries, it is most convenient to begin by finding the
513 Kapitza or interfacial thermal resistance for a thin spherical shell,
514 \begin{equation}
515 R_K = \frac{1}{G} = \frac{\Delta
516 T}{J_r}
517 \end{equation}
518 where $\Delta T$ is the temperature drop from the interior to the
519 exterior of the shell. For two concentric shells, the kinetic energy
520 flux ($J_r$) is not the same (as the surface areas are not the same),
521 but the heat transfer rate, $q_r = J_r A$ is constant. The thermal
522 resistance of a shell with interior radius $r$ is most conveniently
523 written as
524 \begin{equation}
525 R_K = \frac{1}{q_r} \Delta T 4 \pi r^2.
526 \label{eq:RK}
527 \end{equation}
528
529 \begin{figure}
530 \includegraphics[width=\linewidth]{figures/NP20}
531 \caption{A gold nanoparticle with a radius of 20 \AA$\,$ solvated in
532 TraPPE-UA hexane. A kinetic energy flux is applied between the
533 nanoparticle and an outer shell of solvent to obtain the interfacial
534 thermal conductance, $G$, and the interfacial rotational resistance
535 $\Xi^{rr}$ is determined using an angular momentum flux. }
536 \label{fig:NP20}
537 \end{figure}
538
539 To model the thermal conductance across a wide interface (or multiple
540 interfaces) it is useful to consider concentric shells as resistors
541 wired in series. The resistance of the shells is then additive, and
542 the interfacial thermal conductance is the inverse of the total
543 Kapitza resistance:
544 \begin{equation}
545 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
546 T_i\right) 4 \pi r_i^2
547 \end{equation}
548 This series can be extended for any number of adjacent shells,
549 allowing for the calculation of the interfacial thermal conductance
550 for interfaces of considerable thickness, such as self-assembled
551 ligand monolayers on a metal surface.
552
553 \subsubsection{Interfacial thermal conductance of solvated gold
554 nanoparticles}
555 Calculated interfacial thermal conductance ($G$) values for three
556 sizes of gold nanoparticles and a flat Au(111) surface solvated in
557 TraPPE-UA hexane~\cite{Stocker:2013cl} are shown in Table
558 \ref{table:G}.
559
560 \begin{table}
561 \caption{Calculated interfacial thermal conductance ($G$) values for
562 gold nanoparticles of varying radii solvated in TraPPE-UA
563 hexane. The nanoparticle $G$ values are compared to previous
564 simulation results for a Au(111) interface in TraPPE-UA hexane.}
565 \label{table:G}
566 \begin{tabular}{cc}
567 \hline \hline
568 {Nanoparticle Radius} & {$G$}\\
569 {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
570 20 & {47.1 $\pm$ 5.0} \\
571 30 & {45.4 $\pm$ 1.2} \\
572 40 & {46.5 $\pm$ 2.1} \\
573 \hline
574 Au(111) & {30.2} \\
575 \hline \hline
576 \end{tabular}
577 \end{table}
578
579 The introduction of surface curvature increases the interfacial
580 thermal conductance by a factor of approximately $1.5$ relative to the
581 flat interface, although there is no apparent size-dependence in the
582 $G$ values obtained from our calculations. This is quite different
583 behavior than the size-dependence observed by Lervik \textit{et
584 al.}\cite{Lervik:2009fk} in their NEMD simulations of decane
585 droplets in water. In the liquid/liquid case, the surface tension is
586 strongly dependent on the droplet size, and the interfacial
587 resistivity increases with surface tension. In our case, the surface
588 tension of the solid / liquid interface does not have a strong
589 dependence on the particle radius at these particle sizes.
590
591 Simulations of larger nanoparticles may yet demonstrate limiting $G$
592 values close to the flat Au(111) slab, although any spherical particle
593 will have a significant fraction of the surface atoms in non-111
594 facets. It is not clear at this point whether the increase in thermal
595 conductance for the spherical particles is due to the increased
596 population of undercoordinated surface atoms, or to increased solid
597 angle space available for phonon scattering into the solvent. These
598 two possibilities do open up some interesting avenues for
599 investigation.
600
601 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
602 % INTERFACIAL ROTATIONAL FRICTION
603 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
604 \subsection{Interfacial rotational friction}
605 \label{sec:rotation}
606 The interfacial rotational resistance tensor, $\Xi^{rr}$, can be
607 calculated for heterogeneous nanostructure/solvent systems by applying
608 an angular momentum flux between the solvated nanostructure and a
609 spherical shell of solvent at the outer edge of the cluster. An
610 angular velocity gradient develops in response to the applied flux,
611 causing the nanostructure and solvent shell to rotate in opposite
612 directions about a given axis.
613
614 \begin{figure}
615 \includegraphics[width=\linewidth]{figures/E25-75}
616 \caption{A gold prolate ellipsoid of length 65 \AA$\,$ and width 25
617 \AA$\,$ solvated by TraPPE-UA hexane. An angular momentum flux is
618 applied between the ellipsoid and an outer shell of solvent.}
619 \label{fig:E25-75}
620 \end{figure}
621
622 Analytical solutions for the diagonal elements of the rotational
623 resistance tensor for solvated spherical body of radius $r$ under
624 ideal stick boundary conditions can be estimated using Stokes' law
625 \begin{equation}
626 \Xi^{rr}_{stick} = 8 \pi \eta r^3,
627 \label{eq:Xisphere}
628 \end{equation}
629 where $\eta$ is the dynamic viscosity of the surrounding solvent.
630
631 For general ellipsoids with semiaxes $\alpha$, $\beta$, and $\gamma$,
632 Perrin's extension of Stokes' law provides exact solutions for
633 symmetric prolate $(\alpha \geq \beta = \gamma)$ and oblate $(\alpha <
634 \beta = \gamma)$ ellipsoids under ideal stick conditions. The Perrin
635 elliptic integral parameter,
636 \begin{equation}
637 S = \frac{2}{\sqrt{\alpha^2 - \beta^2}} \ln \left[ \frac{\alpha + \sqrt{\alpha^2 - \beta^2}}{\beta} \right].
638 \label{eq:S}
639 \end{equation}
640
641 For a prolate ellipsoidal nanoparticle (see Fig. \ref{fig:E25-75}),
642 the rotational resistance tensor $\Xi^{rr}$ is a $3 \times 3$ diagonal
643 matrix with elements
644 \begin{align}
645 \Xi^{rr}_\alpha =& \frac{32 \pi}{3} \eta \frac{ \left(
646 \alpha^2 - \beta^2 \right) \beta^2}{2\alpha - \beta^2 S}
647 \\ \nonumber \\
648 \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},
649 \label{eq:Xirr}
650 \end{align}
651 corresponding to rotation about the long axis ($\alpha$), and each of
652 the equivalent short axes ($\beta$ and $\gamma$), respectively.
653
654 Previous VSS-RNEMD simulations of the interfacial friction of the
655 planar Au(111) / hexane interface have shown that the flat interface
656 exists within slip boundary conditions.\cite{Kuang:2012fe} Hu and
657 Zwanzig\cite{Zwanzig} investigated the rotational friction
658 coefficients for spheroids under slip boundary conditions and obtained
659 numerial results for a scaling factor to be applied to
660 $\Xi^{rr}_{\mathit{stick}}$ as a function of ratio of the shorter
661 semiaxes to the longer. Under slip conditions, rotation of any
662 sphere, and rotation of a prolate ellipsoid about its long axis does
663 not displace any solvent, so the resulting $\Xi^{rr}_{\mathit{slip}}$
664 approaches $0$. For rotation of a prolate ellipsoid (with the aspect
665 ratio shown here) about its short axis, Hu and Zwanzig obtained
666 $\Xi^{rr}_{\mathit{slip}}$ values of $35.9\%$ of the analytical
667 $\Xi^{rr}_{\mathit{stick}}$ result. This reduced rotational
668 resistance accounts for the reduced interfacial friction under slip
669 boundary conditions.
670
671 The measured rotational friction coefficient,
672 $\Xi^{rr}_{\mathit{meas}}$ at the interface can be extracted from
673 non-periodic VSS-RNEMD simulations quite easily using the total
674 applied torque ($\tau$) and the observed angular velocity of the solid
675 structure ($\omega_\mathrm{solid}$),
676 \begin{equation}
677 \Xi^{rr}_{\mathit{meas}} = \frac{\tau}{\omega_\mathrm{solid}}
678 \label{eq:Xieff}
679 \end{equation}
680
681 The total applied torque required to overcome the interfacial friction
682 and maintain constant rotation of the gold is
683 \begin{equation}
684 \tau = \frac{j_r(\mathbf{L}) \cdot A}{2}
685 \label{eq:tau}
686 \end{equation}
687 where $j_r(\mathbf{L})$ is the applied angular momentum flux, $A$ is
688 the surface area of the solid nanoparticle, and the factor of $2$ is
689 present because the torque is exerted on both the particle and the
690 counter-rotating solvent shell.
691
692 \subsubsection{Rotational friction for gold nanostructures in hexane}
693
694 Table \ref{table:couple} shows the calculated rotational friction
695 coefficients $\Xi^{rr}$ for spherical gold nanoparticles and a prolate
696 ellipsoidal gold nanorod solvated in TraPPE-UA hexane. An angular
697 momentum flux was applied between an outer shell of solvent that was
698 at least 20 \AA\ away from the surface of the particle (region $A$)
699 and the gold nanostructure (region $B$). Dynamic viscosity estimates
700 $(\eta)$ for TraPPE-UA hexane under these particular temperature and
701 pressure conditions was determined by applying a traditional VSS-RNEMD
702 linear momentum flux to a periodic box of this solvent.
703
704 \begin{table}
705 \caption{Comparison of rotational friction coefficients under ideal stick
706 ($\Xi^{rr}_{\mathit{stick}}$) and
707 slip ($\Xi^{rr}_{\mathit{slip}}$) conditions, along with the measured
708 ($\Xi^{rr}_{\mathit{meas}}$) rotational friction coefficients of gold
709 nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid
710 is oriented with the long axis along the $z$ direction.}
711 \label{table:couple}
712 \begin{tabular}{lccccc}
713 \\ \hline \hline
714 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{meas}}$} & {$\Xi^{rr}_{\mathit{meas}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
715 {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
716 Sphere (r = 20 \AA) & {$x = y = z$} & {3314} & 0 & {2386 $\pm$ 14} & {0.720 $\pm$ 0.004}\\
717 Sphere (r = 30 \AA) & {$x = y = z$} & {11749}& 0 & {8415 $\pm$ 274} & {0.716 $\pm$ 0.023}\\
718 Sphere (r = 40 \AA) & {$x = y = z$} & {34464}& 0 & {47544 $\pm$ 3051} & {1.380 $\pm$ 0.089}\\
719 Prolate Ellipsoid & {$x = y$} & {4991} & {1792} & {3128 $\pm$ 166} & {0.627 $\pm$ 0.033}\\
720 Prolate Ellipsoid & {$z$} & {1993} & 0 & {1590 $\pm$ 30} & {0.798 $\pm$ 0.015}
721 \\ \hline \hline
722 \end{tabular}
723 \end{table}
724
725 The results for $\Xi^{rr}_{\mathit{meas}}$ show that, in contrast with
726 the flat Au(111) / hexane interface, gold nanostructures solvated by
727 hexane are closer to stick than slip boundary conditions. At these
728 length scales, the nanostructures are not perfect spheroids due to
729 atomic `roughening' of the surface and therefore experience increased
730 interfacial friction, deviating significantly from the ideal slip
731 case. The 20 and 30~\AA\ radius nanoparticles experience approximately
732 70\% of the ideal stick boundary interfacial friction. Rotation of the
733 ellipsoid about its long axis more closely approaches the stick limit
734 than rotation about the short axis, which at first seems
735 counterintuitive. However, the `propellor' motion caused by rotation
736 around the short axis may exclude solvent from the rotation cavity or
737 move a sufficient amount of solvent along with the gold that a smaller
738 interfacial friction is actually experienced. The largest nanoparticle
739 (40 \AA$\,$ radius) appears to experience interfacial friction in
740 excess of the ideal stick conditions. Although the solvent velocity
741 field does not appear in the calculation of $\Xi^{rr}$, we do note
742 that the largest particle has a solvent velocity field that requires
743 fitting to the $v(r,\theta) = \sin\theta \left( \frac{A}{r} + B r^2
744 \right)$ whereas the solvent velocity fields for the other particles
745 can be fit with just the first term. CHECK THIS.
746
747 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
748 % **DISCUSSION**
749 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
750 \section{Discussion}
751
752 We have adapted the VSS-RNEMD methodology so that it can be used to
753 apply kinetic energy and angular momentum fluxes in explicitly
754 non-periodic simulation geometries. Non-periodic VSS-RNEMD preserves
755 the strengths of the original periodic variant, specifically Boltzmann
756 velocity distributions and minimal thermal anisotropy, while extending
757 the constraints to conserve total energy and total \emph{angular}
758 momentum. This method also maintains the ability to impose the kinetic
759 energy and angular momentum fluxes either jointly or individually.
760
761 This method enables steady-state calculation of interfacial thermal
762 conductance and interfacial rotational friction in heterogeneous
763 clusters. We have demonstrated the abilities of the method on some
764 familiar systems (nanocrystalline gold, SPC/E water) as well as on
765 interfaces that have been studied only in planar geometries. For the
766 bulk systems that are well-understood in periodic geometries, the
767 nonperiodic VSS-RNEMD provides very similar estimates of the thermal
768 conductivity to other simulation methods.
769
770 For nanoparticle-to-solvent heat transfer, we have observed that the
771 interfacial conductance exhibits a $1.5\times$ increase over the
772 planar Au(111) interface, although we do not see any dependence of the
773 conductance on the particle size. Because the surface tension effects
774 present in liquid/liquid heat transfer are not strong contributors
775 here, we suggest two possible mechanisms for the $1.5\times$ increase
776 over the planar surface: (1) the nanoparticles have an increased
777 population of undercoordinated surface atoms that are more efficient
778 at transferring vibrational energy to the solvent, or (2) there is
779 increased solid angle space available for phonon scattering into the
780 solvent. The second of these explanations would have a significant
781 dependence on the radius of spherical particles, although crystalline
782 faceting of the particles may be enough to reduce this dependence on
783 particle radius. These two possibilities do open up some interesting
784 avenues for further exploration.
785
786 One area that this method opens up that was not available for periodic
787 simulation cells is direct computation of the solid/liquid rotational
788 friction coefficients of nanostructures. The systems we have
789 investigated (gold nanospheres and prolate ellipsoids) have analytic
790 stick and approximate slip solutions provided via hydrodynamics, and
791 our molecular simulations indicate that the bare metallic
792 nanoparticles are closer to the stick than they are to slip
793 conditions. This is markedly different behavior than we see for the
794 solid / liquid friction at planar Au(111) interfaces, where the
795 solvents experience a nearly-laminar flow over the surface in previous
796 simulations. Schmidt and Skinner previously computed the behavior of
797 spherical tag particles in molecular dynamics simulations, and showed
798 that {\it slip} boundary conditions for translational resistance may
799 be more appropriate for molecule-sized spheres embedded in a sea of
800 spherical solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx} The
801 size of the gold nanoparticles studied here and the structuring of the
802 surfaces of the particles appears to bring the behavior closer to
803 stick boundaries. Exactly where the slip-stick crossover takes place
804 is also an interesting avenue for exploration.
805
806 The ability to interrogate explicitly non-periodic effects
807 (e.g. surface curvature, edge effects between facets, and the splay of
808 ligand surface groups) on interfacial transport means that this method
809 can be applied to systems of broad experimental and theoretical
810 interest.
811
812 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
813 % **ACKNOWLEDGMENTS**
814 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
815 \begin{acknowledgement}
816 The authors thank Dr. Shenyu Kuang for helpful discussions. Support
817 for this project was provided by the National Science Foundation
818 under grant CHE-0848243. Computational time was provided by the
819 Center for Research Computing (CRC) at the University of Notre Dame.
820 \end{acknowledgement}
821
822
823 \newpage
824
825 \bibliography{nonperiodicVSS}
826
827 %\end{doublespace}
828 \end{document}

Properties

Name Value
svn:executable