ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4072
Committed: Thu Mar 13 21:48:39 2014 UTC (10 years, 6 months ago) by kstocke1
Content type: application/x-tex
File size: 36697 byte(s)
Log Message:

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}
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, however, the RNEMD methods have only been used in periodic
140 simulation cells where the exchange regions are physically separated
141 along one of the axes of the simulation cell. This limits the
142 applicability to infinite planar interfaces which are perpendicular to
143 the applied flux. In order to model steady-state non-equilibrium
144 distributions for curved surfaces (e.g. hot nanoparticles in contact
145 with colder solvent), or for regions that are not planar slabs, the
146 method requires some generalization for non-parallel exchange regions.
147 In the following sections, we present a new velocity shearing and
148 scaling (VSS) RNEMD algorithm which has been explicitly designed for
149 non-periodic simulations, and use the method to compute some thermal
150 transport and solid-liquid friction at the surfaces of spherical and
151 ellipsoidal nanoparticles.
152
153 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
154 % **METHODOLOGY**
155 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156 \section{Velocity shearing and scaling (VSS) for non-periodic systems}
157
158 The original periodic VSS-RNEMD approach uses a series of simultaneous
159 velocity shearing and scaling exchanges between the two
160 slabs.\cite{Kuang:2012fe} This method imposes energy and linear
161 momentum conservation constraints while simultaneously creating a
162 desired flux between the two slabs. These constraints ensure that all
163 configurations are sampled from the same microcanonical (NVE)
164 ensemble.
165
166 \begin{figure}
167 \includegraphics[width=\linewidth]{figures/npVSS2}
168 \caption{Schematics of periodic (left) and non-periodic (right)
169 Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
170 flux is applied from region B to region A. Thermal gradients are
171 depicted by a color gradient. Linear or angular velocity gradients
172 are shown as arrows.}
173 \label{fig:VSS}
174 \end{figure}
175
176 We have extended the VSS method for use in {\it non-periodic}
177 simulations, in which the ``slabs'' have been generalized to two
178 separated regions of space. These regions could be defined as
179 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
180 can be defined in terms of a dynamically changing ``hull'' comprising
181 the surface atoms of the cluster. This latter definition is identical
182 to the hull used in the Langevin Hull algorithm.\cite{Vardeman2011}
183 For the non-periodic variant, the constraints fix both the total
184 energy and total {\it angular} momentum of the system while
185 simultaneously imposing a thermal and angular momentum flux between
186 the two regions.
187
188 After a time interval of $\Delta t$, the particle velocities
189 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
190 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
191 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$). The
192 scalars $a$ and $b$ collectively provide a thermal exchange between
193 the two regions. One of the values is larger than 1, and the other
194 smaller. To conserve total energy and angular momentum, the values of
195 these two scalars are coupled. The vectors ($\mathbf{c}_a$ and
196 $\mathbf{c}_b$) provide a relative rotational shear to the velocities
197 of the particles within the two regions, and these vectors must also
198 be coupled to constrain the total angular momentum.
199
200 Once the values of the scaling and shearing factors are known, the
201 velocity changes are applied,
202 \begin{displaymath}
203 \begin{array}{rclcl}
204 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
205 \underline{\mathrm{rotational~shearing}} \\ \\
206 \mathbf{v}_i $~~~$\leftarrow &
207 a \left(\mathbf{v}_i - \langle \omega_a
208 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
209 \mathbf{v}_j $~~~$\leftarrow &
210 b \left(\mathbf{v}_j - \langle \omega_b
211 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
212 \end{array}
213 \end{displaymath}
214 Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
215 velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
216 (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
217 to their velocities. The amount of shear is governed by the imposed angular momentum flux,
218 $\mathbf{j}_r(\mathbf{L})$,
219 \begin{eqnarray}
220 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
221 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
222 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
223 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
224 \end{eqnarray}
225 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia
226 tensor for each of the two shells.
227
228 To simultaneously impose a thermal flux ($J_r$) between the shells we
229 use energy conservation constraints,
230 \begin{eqnarray}
231 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
232 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
233 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
234 \cdot \mathbf{c}_a \label{eq:Kc}\\
235 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
236 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
237 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
238 \end{eqnarray}
239 Simultaneous solution of these quadratic formulae for the scaling coefficients, $a$ and $b$, will ensure that
240 the simulation samples from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is the instantaneous
241 translational kinetic energy of each shell. At each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
242 $\mathbf{c}_b$, subject to the imposed angular momentum flux, $j_r(\mathbf{L})$, and thermal flux, $J_r$,
243 values. The new particle velocities are computed, and the simulation continues. System configurations after the
244 transformations have exactly the same energy ({\it and} angular momentum) as before the moves.
245
246 As the simulation progresses, the velocity transformations can be
247 performed on a regular basis, and the system will develop a
248 temperature and/or angular velocity gradient in response to the
249 applied flux. By fitting the radial temperature gradient, it is
250 straightforward to obtain the bulk thermal conductivity,
251 \begin{equation}
252 J_r = -\lambda \left( \frac{\partial T}{\partial r}\right)
253 \end{equation}
254 from the radial kinetic energy flux $(J_r)$ that was applied during
255 the simulation. In practice, it is significantly easier to use the
256 integrated form of Fourier's law for spherical geometries. In
257 sections \ref{sec:thermal} -- \ref{sec:rotation} we outline ways of
258 obtaining interfacial transport coefficients from these RNEMD
259 simulations.
260
261 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262 % **COMPUTATIONAL DETAILS**
263 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264 \section{Computational Details}
265
266 The new VSS-RNEMD methodology for non-periodic system geometries has
267 been implemented in our group molecular dynamics code,
268 OpenMD.\cite{Meineke:2005gd,openmd} We have tested the new method to
269 calculate the thermal conductance of a gold nanoparticle and SPC/E
270 water cluster, and compared the results with previous bulk RNEMD
271 values, as well as experiment. We have also investigated the
272 interfacial thermal conductance and interfacial rotational friction
273 for gold nanostructures solvated in hexane as a function of
274 nanoparticle size and shape.
275
276 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277 % FORCE FIELD PARAMETERS
278 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
279 \subsection{Force field}
280
281 Gold -- gold interactions are described by the quantum Sutton-Chen
282 (QSC) model.\cite{PhysRevB.59.3527} The QSC parameters are tuned to
283 experimental properties such as density, cohesive energy, and elastic
284 moduli and include zero-point quantum corrections.
285
286 The SPC/E water model~\cite{Berendsen87} is particularly useful for
287 validation of conductivities and shear viscosities. This model has
288 been used to previously test other RNEMD and NEMD approaches, and
289 there are reported values for thermal conductivies and shear
290 viscosities at a wide range of thermodynamic conditions that are
291 available for direct comparison.\cite{Bedrov:2000,Kuang:2010if}
292
293 Hexane molecules are described by the TraPPE united atom
294 model,\cite{TraPPE-UA.alkanes} which provides good computational
295 efficiency and reasonable accuracy for bulk thermal conductivity
296 values. In this model, sites are located at the carbon centers for
297 alkyl groups. Bonding interactions, including bond stretches and bends
298 and torsions, were used for intra-molecular sites closer than 3
299 bonds. For non-bonded interactions, Lennard-Jones potentials were
300 used. We have previously utilized both united atom (UA) and all-atom
301 (AA) force fields for thermal
302 conductivity,\cite{Kuang:2011ef,Kuang:2012fe,Stocker:2013cl} and since
303 the united atom force fields cannot populate the high-frequency modes
304 that are present in AA force fields, they appear to work better for
305 modeling thermal conductance at metal/ligand interfaces.
306
307 Gold -- hexane nonbonded interactions are governed by pairwise
308 Lennard-Jones parameters derived from Vlugt \emph{et
309 al}.\cite{vlugt:cpc2007154} They fitted parameters for the
310 interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the
311 effective potential of Hautman and Klein for the Au(111)
312 surface.\cite{hautman:4994}
313
314 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
315 % NON-PERIODIC DYNAMICS
316 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
317 % \subsection{Dynamics for non-periodic systems}
318 %
319 % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
320 % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
321 % compressibilities, which are typically problematic for traditional affine transform methods. We have had
322 % success applying this method to several different systems including bare metal nanoparticles, liquid water
323 % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
324 % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
325 % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
326 % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
327 % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
328 % subject to standard Newtonian dynamics.
329
330 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
331 % SIMULATION PROTOCOL
332 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
333 \subsection{Simulation protocol}
334
335 In all cases, systems were equilibrated under non-periodic
336 isobaric-isothermal (NPT) conditions -- using the Langevin Hull
337 methodology\cite{Vardeman2011} -- before any non-equilibrium methods
338 were introduced. For heterogeneous systems, the gold nanoparticles and
339 ellipsoids were created from a bulk fcc lattice and were thermally
340 equilibrated before being solvated in hexane. Packmol\cite{packmol}
341 was used to solvate previously equilibrated gold nanostructures within
342 a spherical droplet of hexane.
343
344 Once equilibrated, thermal or angular momentum fluxes were applied for
345 1 - 2 ns, until stable temperature or angular velocity gradients had
346 developed. Systems containing liquids were run under moderate pressure
347 (5 atm) and temperatures (230 K) to avoid the formation of a vapor
348 layer at the boundary of the cluster. Pressure was applied to the
349 system via the non-periodic Langevin Hull.\cite{Vardeman2011} However,
350 thermal coupling to the external temperature and pressure bath was
351 removed to avoid interference with the imposed RNEMD flux.
352
353 Because the method conserves \emph{total} angular momentum, systems
354 which contain a metal nanoparticle embedded in a significant volume of
355 solvent will still experience nanoparticle diffusion inside the
356 solvent droplet. To aid in computing the rotational friction in these
357 systems, a single gold atom at the origin of the coordinate system was
358 assigned a mass $10,000 \times$ its original mass. The bonded and
359 nonbonded interactions for this atom remain unchanged and the heavy
360 atom is excluded from the RNEMD exchanges. The only effect of this
361 gold atom is to effectively pin the nanoparticle at the origin of the
362 coordinate system, while still allowing for rotation. For rotation of
363 the gold ellipsoids we added two of these heavy atoms along the axis
364 of rotation, separated by an equal distance from the origin of the
365 coordinate system. These heavy atoms prevent off-axis tumbling of the
366 nanoparticle and allow for measurement of rotational friction relative
367 to a particular axis of the ellipsoid.
368
369 Angular velocity data was collected for the heterogeneous systems
370 after a brief period of imposed flux to initialize rotation of the
371 solvated nanostructure. Doing so ensures that we overcome the initial
372 static friction and calculate only the \emph{dynamic} interfacial
373 rotational friction.
374
375 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
376 % THERMAL CONDUCTIVITIES
377 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378 \subsection{Thermal conductivities}
379 \label{sec:thermal}
380
381 To compute the thermal conductivities of bulk materials, the
382 integrated form of Fourier's Law of heat conduction in radial
383 coordinates can be used to obtain an expression for the heat transfer
384 rate between concentric spherical shells:
385 \begin{equation}
386 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
387 \label{eq:Q}
388 \end{equation}
389 The heat transfer rate, $q_r$, is constant in spherical geometries,
390 while the heat flux, $J_r$ depends on the surface area of the two
391 shells. $\lambda$ is the thermal conductivity, and $T_{a,b}$ and
392 $r_{a,b}$ are the temperatures and radii of the two concentric RNEMD
393 regions, respectively.
394
395 A kinetic energy flux is created using VSS-RNEMD moves, and the
396 temperature in each of the radial shells is recorded. The resulting
397 temperature profiles are analyzed to yield information about the
398 interfacial thermal conductance. As the simulation progresses, the
399 VSS moves are performed on a regular basis, and the system develops a
400 thermal or velocity gradient in response to the applied flux. Once a
401 stable thermal gradient has been established between the two regions,
402 the thermal conductivity, $\lambda$, can be calculated using a linear
403 regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
404
405 \begin{equation}
406 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
407 \label{eq:lambda}
408 \end{equation}
409
410 The rate of heat transfer, $q_r$, between the two RNEMD regions is
411 easily obtained from either the applied kinetic energy flux and the
412 area of the smaller of the two regions, or from the total amount of
413 transferred kinetic energy and the run time of the simulation.
414 \begin{equation}
415 q_r = J_r A = \frac{KE}{t}
416 \label{eq:heat}
417 \end{equation}
418
419 \subsubsection{Thermal conductivity of nanocrystalline gold}
420 Calculated values for the thermal conductivity of a 40 \AA$~$ radius
421 gold nanoparticle (15707 atoms) at a range of kinetic energy flux
422 values are shown in Table \ref{table:goldTC}. For these calculations,
423 the hot and cold slabs were excluded from the linear regression of the
424 thermal gradient.
425
426 \begin{longtable}{ccc}
427 \caption{Calculated thermal conductivity of a crystalline gold
428 nanoparticle of radius 40 \AA. Calculations were performed at 300
429 K and ambient density. Gold-gold interactions are described by the
430 Quantum Sutton-Chen potential.}
431 \\ \hline \hline
432 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
433 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
434 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
435 6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
436 1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
437 3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
438 6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
439 \hline
440 This work & & 1.004 $\pm$ 0.009
441 \\ \hline \hline
442 \label{table:goldTC}
443 \end{longtable}
444
445 The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly
446 dependent on the applied kinetic energy flux $J_r$. Calculated thermal
447 conductivity values compare well with previous bulk QSC values of 1.08
448 -- 1.26 W / m $\cdot$ K\cite{Kuang:2010if}, though still significantly
449 lower than the experimental value of 320 W / m $\cdot$ K, as the QSC
450 force field neglects significant electronic contributions to heat
451 conduction.
452
453 \subsubsection{Thermal conductivity of a droplet of SPC/E water}
454
455 Calculated values for the thermal conductivity of a cluster of 6912
456 SPC/E water molecules are shown in Table \ref{table:waterTC}. As with
457 the gold nanoparticle thermal conductivity calculations, the RNEMD
458 regions were excluded from the $\langle \frac{dT}{dr} \rangle$ fit.
459
460 \begin{longtable}{ccc}
461 \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E
462 water molecules. Calculations were performed at 300 K and 5 atm.}
463 \\ \hline \hline
464 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
465 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
466 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
467 3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
468 6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
469 \hline
470 This work & & 0.884 $\pm$ 0.013\\
471 Zhang, et al\cite{Zhang2005} & & 0.81 \\
472 R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
473 Experiment\cite{WagnerKruse} & & 0.61
474 \\ \hline \hline
475 \label{table:waterTC}
476 \end{longtable}
477
478 Again, the measured slope is linearly dependent on the applied kinetic
479 energy flux $J_r$. The average calculated thermal conductivity from
480 this work, $0.8841$ W / m $\cdot$ K, compares very well to previous
481 non-equilibrium molecular dynamics results\cite{Romer2012, Zhang2005}
482 and experimental values.\cite{WagnerKruse}
483
484 \begin{figure}
485 \includegraphics[width=\linewidth]{figures/lambda_G}
486 \caption{}
487 \label{fig:lambda_G}
488 \end{figure}
489
490
491 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
492 % INTERFACIAL THERMAL CONDUCTANCE
493 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
494 \subsection{Interfacial thermal conductance}
495 \label{sec:interfacial}
496
497 The interfacial thermal conductance, $G$, of a heterogeneous interface
498 located at $r_0$ can be understood as the change in thermal
499 conductivity in a direction normal $(\mathbf{\hat{n}})$ to the
500 interface,
501 \begin{equation}
502 G = \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{r_0}
503 \end{equation}
504 For heterogeneous systems such as the solvated nanoparticle shown in
505 Figure \ref{fig:NP20}, the interfacial thermal conductance at the
506 surface of the nanoparticle can be determined using a kinetic energy
507 flux applied using the RNEMD method developed above. It is most
508 convenient to compute the Kaptiza or interfacial thermal resistance
509 for a thin spherical shell,
510 \begin{equation}
511 R_K = \frac{1}{G} = \frac{\Delta
512 T}{J_r}
513 \end{equation}
514 where $\Delta T$ is the temperature drop from the interior to the
515 exterior of the shell. For two neighboring shells, the kinetic energy
516 flux ($J_r$) is not the same (as the surface areas are not the same),
517 but the heat transfer rate, $q_r = J_r A$ is constant. The thermal
518 resistance of a shell with interior radius $r$ is most conveniently
519 written as
520 \begin{equation}
521 R_K = \frac{1}{q_r} \Delta T 4 \pi r^2.
522 \label{eq:RK}
523 \end{equation}
524
525
526 \begin{figure}
527 \includegraphics[width=\linewidth]{figures/NP20}
528 \caption{A gold nanoparticle with a radius of 20 \AA$\,$ solvated in
529 TraPPE-UA hexane. A kinetic energy flux is applied between the
530 nanoparticle and an outer shell of solvent to obtain the interfacial
531 thermal conductance, $G$, and the interfacial rotational resistance
532 $\Xi^{rr}$ is determined using an angular momentum flux. }
533 \label{fig:NP20}
534 \end{figure}
535
536 To model the thermal conductance across an interface (or multiple
537 interfaces) it is useful to consider the shells as resistors wired in
538 series. The resistance of the shells is then additive, and the
539 interfacial thermal conductance is the inverse of the total Kapitza
540 resistance:
541 \begin{equation}
542 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
543 T_i\right) 4 \pi r_i^2
544 \end{equation}
545 This series can be expanded for any number of adjacent shells,
546 allowing for the calculation of the interfacial thermal conductance
547 for interfaces of considerable thickness, such as self-assembled
548 ligand monolayers on a metal surface.
549
550 \subsubsection{Interfacial thermal conductance of solvated gold
551 nanoparticles}
552 Calculated interfacial thermal conductance ($G$) values for three
553 sizes of gold nanoparticles and a flat Au(111) surface solvated in
554 TraPPE-UA hexane are shown in Table \ref{table:G}.
555
556 \begin{longtable}{ccc}
557 \caption{Calculated interfacial thermal conductance ($G$) values for
558 gold nanoparticles of varying radii solvated in TraPPE-UA
559 hexane. The nanoparticle $G$ values are compared to previous
560 simulation results for a Au(111) interface in TraPPE-UA hexane.}
561 \\ \hline \hline
562 {Nanoparticle Radius} & {$G$}\\
563 {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
564 20 & {47.1 $\pm$ 5.0} \\
565 30 & {45.4 $\pm$ 1.2} \\
566 40 & {46.5 $\pm$ 2.1} \\
567 \hline
568 Au(111) & {30.2}
569 \\ \hline \hline
570 \label{table:G}
571 \end{longtable}
572
573 The introduction of surface curvature increases the interfacial
574 thermal conductance by a factor of approximately $1.5$ relative to the
575 flat interface. There are no significant differences in the $G$ values
576 for the varying nanoparticle sizes. It seems likely that for the range
577 of nanoparticle sizes represented here, any particle size effects are
578 not evident. Simulations of larger nanoparticles may yet demonstrate
579 a limiting $G$ value close tothe flat Au(111) slab but these would
580 require prohibitively costly numbers of atoms.
581
582 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
583 % INTERFACIAL ROTATIONAL FRICTION
584 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
585 \subsection{Interfacial rotational friction}
586 \label{sec:rotation}
587 The interfacial rotational resistance tensor, $\Xi^{rr}$, can be
588 calculated for heterogeneous nanostructure/solvent systems by applying
589 an angular momentum flux between the solvated nanostructure and a
590 spherical shell of solvent at the outer edge of the cluster. An
591 angular velocity gradient develops in response to the applied flux,
592 causing the nanostructure and solvent shell to rotate in opposite
593 directions about a given axis.
594
595 \begin{figure}
596 \includegraphics[width=\linewidth]{figures/E25-75}
597 \caption{A gold prolate ellipsoid of length 65 \AA$\,$ and width 25
598 \AA$\,$ solvated by TraPPE-UA hexane. An angular momentum flux is
599 applied between the ellipsoid and an outer shell of solvent.}
600 \label{fig:E25-75}
601 \end{figure}
602
603 Analytical solutions for the diagonal elements of the rotational
604 resistance tensor for solvated spherical body of radius $r$ under
605 ideal stick boundary conditions can be estimated using Stokes' law
606 \begin{equation}
607 \Xi^{rr}_{stick} = 8 \pi \eta r^3,
608 \label{eq:Xisphere}
609 \end{equation}
610 where $\eta$ is the dynamic viscosity of the surrounding solvent.
611
612 For general ellipsoids with semiaxes $\alpha$, $\beta$, and $\gamma$,
613 Perrin's extension of Stokes' law provides exact solutions for
614 symmetric prolate $(\alpha \geq \beta = \gamma)$ and oblate $(\alpha <
615 \beta = \gamma)$ ellipsoids under ideal stick conditions. For
616 simplicity, we define the Perrin factor, $S$,
617
618 \begin{equation}
619 S = \frac{2}{\sqrt{\alpha^2 - \beta^2}} \ln \left[ \frac{\alpha + \sqrt{\alpha^2 - \beta^2}}{\beta} \right].
620 \label{eq:S}
621 \end{equation}
622
623 For a prolate ellipsoidal nanoparticle (see Fig. \ref{fig:E25-75}),
624 the rotational resistance tensor $\Xi^{rr}$ is a $3 \times 3$ diagonal
625 matrix with elements
626 \begin{align}
627 \Xi^{rr}_\alpha =& \frac{32 \pi}{3} \eta \frac{ \left(
628 \alpha^2 - \beta^2 \right) \beta^2}{2\alpha - \beta^2 S}
629 \\ \nonumber \\
630 \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},
631 \label{eq:Xirr}
632 \end{align}
633 corresponding to rotation about the long axis ($\alpha$), and each of
634 the equivalent short axes ($\beta$ and $\gamma$), respectively.
635
636 Previous VSS-RNEMD simulations of the interfacial friction of the
637 planar Au(111) / hexane interface have shown that the interface exists
638 within slip boundary conditions.\cite{Kuang:2012fe} Hu and
639 Zwanzig\cite{Zwanzig} investigated the rotational friction
640 coefficients for spheroids under slip boundary conditions and obtained
641 numerial results for a scaling factor to be applied to
642 $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau = \frac{\beta,\gamma}{\alpha}$, the ratio of the
643 shorter semiaxes and the longer semiaxis of the spheroid. For the
644 sphere and prolate ellipsoid shown here, the values of $\tau$ are $1$
645 and $0.3939$, respectively. Under ``slip'' conditions,
646 $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate
647 ellipsoid about its long axis approaches $0$, as no solvent is
648 displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for
649 rotation of the prolate ellipsoid about its short axis is $35.9\%$ of
650 the analytical $\Xi^{rr}_{\mathit{stick}}$ result, accounting for the
651 reduced interfacial friction under ``slip'' boundary conditions.
652
653
654 An
655 $\eta$ value for TraPPE-UA hexane under these particular temperature
656 and pressure conditions was determined by applying a traditional
657 VSS-RNEMD linear momentum flux to a periodic box of solvent.
658
659 The effective rotational friction coefficient, $\Xi^{rr}_{\mathit{eff}}$ at the interface can be extracted from non-periodic VSS-RNEMD simulations quite easily using the applied torque ($\tau$) and the observed angular velocity of the gold structure ($\omega_{Au}$)
660
661 \begin{equation}
662 \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
663 \label{eq:Xieff}
664 \end{equation}
665
666 The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
667
668 \begin{equation}
669 \tau = \frac{L}{2 t}
670 \label{eq:tau}
671 \end{equation}
672
673 where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
674
675
676 Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold
677 nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied
678 between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius,
679 respectively.
680
681 \begin{longtable}{lccccc}
682 \caption{Comparison of rotational friction coefficients under ideal ``slip'' ($\Xi^{rr}_{\mathit{slip}}$) and ``stick'' ($\Xi^{rr}_{\mathit{stick}}$) conditions and effective ($\Xi^{rr}_{\mathit{eff}}$) rotational friction coefficients of gold nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid is oriented with the long axis along the $z$ direction.}
683 \\ \hline \hline
684 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
685 {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
686 Sphere (r = 20 \AA) & {$x = y = z$} & 0 & {2386 $\pm$ 14} & {3314} & {0.720}\\
687 Sphere (r = 30 \AA) & {$x = y = z$} & 0 & {8415 $\pm$ 274} & {11749} & {0.716}\\
688 Sphere (r = 40 \AA) & {$x = y = z$} & 0 & {47544 $\pm$ 3051} & {34464} & {1.380}\\
689 Prolate Ellipsoid & {$x = y$} & {1792} & {3128 $\pm$ 166} & {4991} & {0.627}\\
690 Prolate Ellipsoid & {$z$} & 0 & {1590 $\pm$ 30} & {1993} & {0.798}
691 \\ \hline \hline
692 \label{table:couple}
693 \end{longtable}
694
695 The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold
696 structures solvated by hexane do not exist in the ``slip'' boundary condition. At this length scale, the
697 nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience
698 increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius
699 nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of
700 the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short
701 axis, which at first seems counterintuitive. However, the `propellor' motion caused by rotation about the
702 short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the
703 gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius)
704 appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of
705 surface features or anomalous solvent behaviors that are not fully understood at this time.
706
707 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
708 % **DISCUSSION**
709 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710 \section{Discussion}
711
712 We have demonstrated a novel adaptation of the VSS-RNEMD methodology for the application of thermal and angular momentum fluxes in explicitly non-periodic system geometries. Non-periodic VSS-RNEMD preserves the virtues of traditional VSS-RNEMD, namely Boltzmann thermal velocity distributions and minimal thermal anisotropy, while extending the constraints to conserve total energy and total \emph{angular} momentum. We also still have the ability to impose the thermal and angular momentum fluxes simultaneously or individually.
713
714 Most strikingly, this method enables calculation of thermal conductivity in homogeneous non-periodic geometries, as well as interfacial thermal conductance and interfacial rotational friction in heterogeneous clusters. The ability to interrogate explicitly non-periodic effects -- such as surface curvature -- on interfacial transport properties is an exciting prospect that will be explored in the future.
715
716 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
717 % **ACKNOWLEDGMENTS**
718 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
719 \begin{acknowledgement}
720 The authors thank Dr. Shenyu Kuang for helpful discussions. Support
721 for this project was provided by the National Science Foundation
722 under grant CHE-0848243. Computational time was provided by the
723 Center for Research Computing (CRC) at the University of Notre Dame.
724 \end{acknowledgement}
725
726
727 \newpage
728
729 \bibliography{nonperiodicVSS}
730
731 %\end{doublespace}
732 \end{document}

Properties

Name Value
svn:executable