ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4079
Committed: Fri Mar 14 20:14:14 2014 UTC (10 years, 6 months ago) by kstocke1
Content type: application/x-tex
File size: 41570 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=3.2cm]{figures/npVSS} \includegraphics[width=3.7cm]{figures/NP20}
35 \includegraphics[width=9cm]{figures/TOC}
36 \end{tocentry}
37
38
39 \newcolumntype{A}{p{1.5in}}
40 \newcolumntype{B}{p{0.75in}}
41
42 % \author{Kelsey M. Stocker and J. Daniel
43 % Gezelter\footnote{Corresponding author. \ Electronic mail:
44 % gezelter@nd.edu} \\
45 % 251 Nieuwland Science Hall, \\
46 % Department of Chemistry and Biochemistry,\\
47 % University of Notre Dame\\
48 % Notre Dame, Indiana 46556}
49
50 %\date{\today}
51
52 %\maketitle
53
54 %\begin{doublespace}
55
56 \begin{abstract}
57
58 We present a new reverse non-equilibrium molecular dynamics (RNEMD)
59 method that can be used with non-periodic simulation cells. This
60 method applies thermal and/or angular momentum fluxes between two
61 arbitrary regions of the simulation, and is capable of creating
62 stable temperature and angular velocity gradients while conserving
63 total energy and angular momentum. One particularly useful
64 application is the exchange of kinetic energy between two concentric
65 spherical regions, which can be used to generate thermal transport
66 between nanoparticles and the solvent that surrounds them. The
67 rotational couple to the solvent (a measure of interfacial friction)
68 is also available via this method. As demonstrations and tests of
69 the new method, we have computed the thermal conductivities of gold
70 nanoparticles and water clusters, the shear viscosity of a water
71 cluster, the interfacial thermal conductivity ($G$) of a solvated
72 gold nanoparticle and the interfacial friction of a variety of
73 solvated gold nanostructures.
74
75 \end{abstract}
76
77 \newpage
78
79 %\narrowtext
80
81 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
82 % **INTRODUCTION**
83 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
84 \section{Introduction}
85
86 Non-equilibrium molecular dynamics (NEMD) methods impose a temperature
87 or velocity {\it gradient} on a
88 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}
89 and use linear response theory to connect the resulting thermal or
90 momentum {\it flux} to transport coefficients of bulk materials,
91 \begin{equation}
92 j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}, \hspace{0.5in}
93 J_z = \lambda \frac{\partial T}{\partial z}.
94 \end{equation}
95 Here, $\frac{\partial T}{\partial z}$ and $\frac{\partial
96 v_x}{\partial z}$ are the imposed thermal and momentum gradients,
97 and as long as the imposed gradients are relatively small, the
98 corresponding fluxes, $J_z$ and $j_z(p_x)$, have a linear relationship
99 to the gradients. The coefficients that provide this relationship
100 correspond to physical properties of the bulk material, either the
101 shear viscosity $(\eta)$ or thermal conductivity $(\lambda)$. For
102 systems which include phase boundaries or interfaces, it is often
103 unclear what gradient (or discontinuity) should be imposed at the
104 boundary between materials.
105
106 In contrast, reverse Non-Equilibrium Molecular Dynamics (RNEMD)
107 methods impose an unphysical {\it flux} between different regions or
108 ``slabs'' of the simulation
109 box.\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Patel:2005zm,Shenogina:2009ix,Tenney:2010rp,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
110 The system responds by developing a temperature or velocity {\it
111 gradient} between the two regions. The gradients which develop in
112 response to the applied flux have the same linear response
113 relationships to the transport coefficient of interest. Since the
114 amount of the applied flux is known exactly, and measurement of a
115 gradient is generally less complicated, imposed-flux methods typically
116 take shorter simulation times to obtain converged results. At
117 interfaces, the observed gradients often exhibit near-discontinuities
118 at the boundaries between dissimilar materials. RNEMD methods do not
119 need many trajectories to provide information about transport
120 properties, and they have become widely used to compute thermal and
121 mechanical transport in both homogeneous liquids and
122 solids~\cite{Muller-Plathe:1997wq,Muller-Plathe:1999ao,Tenney:2010rp}
123 as well as heterogeneous
124 interfaces.\cite{Patel:2005zm,Shenogina:2009ix,Kuang:2010if,Kuang:2011ef,Kuang:2012fe,Stocker:2013cl}
125
126 The strengths of specific algorithms for imposing the flux between two
127 different slabs of the simulation cell has been the subject of some
128 renewed interest. The original RNEMD approach used kinetic energy or
129 momentum exchange between particles in the two slabs, either through
130 direct swapping of momentum vectors or via virtual elastic collisions
131 between atoms in the two regions. There have been recent
132 methodological advances which involve scaling all particle velocities
133 in both slabs.\cite{Kuang:2010if,Kuang:2012fe} Constraint equations
134 are simultaneously imposed to require the simulation to conserve both
135 total energy and total linear momentum. The most recent and simplest
136 of the velocity scaling approaches allows for simultaneous shearing
137 (to provide viscosity estimates) as well as scaling (to provide
138 information about thermal conductivity).\cite{Kuang:2012fe}
139
140 To date, the primary method of studying heat transfer at {\it curved}
141 nanoscale interfaes has involved transient non-equilibrium molecular
142 dynamics and temperature relaxation.\cite{Lervik:2009fk} RNEMD methods
143 have only been used in periodic simulation cells where the exchange
144 regions are physically separated along one of the axes of the
145 simulation cell. This limits the applicability to infinite planar
146 interfaces which are perpendicular to the applied flux. In order to
147 model steady-state non-equilibrium distributions for curved surfaces
148 (e.g. hot nanoparticles in contact with colder solvent), or for
149 regions that are not planar slabs, the method requires some
150 generalization for non-parallel exchange regions. In the following
151 sections, we present a new velocity shearing and scaling (VSS) RNEMD
152 algorithm which has been explicitly designed for non-periodic
153 simulations, and use the method to compute some thermal transport and
154 solid-liquid friction at the surfaces of spherical and ellipsoidal
155 nanoparticles.
156
157 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158 % **METHODOLOGY**
159 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160 \section{Velocity shearing and scaling (VSS) for non-periodic systems}
161
162 The original periodic VSS-RNEMD approach uses a series of simultaneous
163 velocity shearing and scaling exchanges between the two
164 slabs.\cite{Kuang:2012fe} This method imposes energy and linear
165 momentum conservation constraints while simultaneously creating a
166 desired flux between the two slabs. These constraints ensure that all
167 configurations are sampled from the same microcanonical (NVE)
168 ensemble.
169
170 \begin{figure}
171 \includegraphics[width=\linewidth]{figures/npVSS2}
172 \caption{Schematics of periodic (left) and non-periodic (right)
173 Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
174 flux is applied from region B to region A. Thermal gradients are
175 depicted by a color gradient. Linear or angular velocity gradients
176 are shown as arrows.}
177 \label{fig:VSS}
178 \end{figure}
179
180 We have extended the VSS method for use in {\it non-periodic}
181 simulations, in which the ``slabs'' have been generalized to two
182 separated regions of space. These regions could be defined as
183 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
184 can be defined in terms of a dynamically changing ``hull'' comprising
185 the surface atoms of the cluster. This latter definition is identical
186 to the hull used in the Langevin Hull algorithm.\cite{Vardeman2011}
187 For the non-periodic variant, the constraints fix both the total
188 energy and total {\it angular} momentum of the system while
189 simultaneously imposing a thermal and angular momentum flux between
190 the two regions.
191
192 After a time interval of $\Delta t$, the particle velocities
193 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
194 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
195 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$). The
196 scalars $a$ and $b$ collectively provide a thermal exchange between
197 the two regions. One of the values is larger than 1, and the other
198 smaller. To conserve total energy and angular momentum, the values of
199 these two scalars are coupled. The vectors ($\mathbf{c}_a$ and
200 $\mathbf{c}_b$) provide a relative rotational shear to the velocities
201 of the particles within the two regions, and these vectors must also
202 be coupled to constrain the total angular momentum.
203
204 Once the values of the scaling and shearing factors are known, the
205 velocity changes are applied,
206 \begin{displaymath}
207 \begin{array}{rclcl}
208 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
209 \underline{\mathrm{rotational~shearing}} \\ \\
210 \mathbf{v}_i $~~~$\leftarrow &
211 a \left(\mathbf{v}_i - \langle \omega_a
212 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
213 \mathbf{v}_j $~~~$\leftarrow &
214 b \left(\mathbf{v}_j - \langle \omega_b
215 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
216 \end{array}
217 \end{displaymath}
218 Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
219 velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
220 (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
221 to their velocities. The amount of shear is governed by the imposed angular momentum flux,
222 $\mathbf{j}_r(\mathbf{L})$,
223 \begin{eqnarray}
224 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
225 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
226 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
227 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
228 \end{eqnarray}
229 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia
230 tensor for each of the two shells.
231
232 To simultaneously impose a thermal flux ($J_r$) between the shells we
233 use energy conservation constraints,
234 \begin{eqnarray}
235 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
236 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
237 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
238 \cdot \mathbf{c}_a \label{eq:Kc}\\
239 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
240 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
241 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
242 \end{eqnarray}
243 Simultaneous solution of these quadratic formulae for the scaling coefficients, $a$ and $b$, will ensure that
244 the simulation samples from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is the instantaneous
245 translational kinetic energy of each shell. At each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
246 $\mathbf{c}_b$, subject to the imposed angular momentum flux, $j_r(\mathbf{L})$, and thermal flux, $J_r$,
247 values. The new particle velocities are computed, and the simulation continues. System configurations after the
248 transformations have exactly the same energy ({\it and} angular momentum) as before the moves.
249
250 As the simulation progresses, the velocity transformations can be
251 performed on a regular basis, and the system will develop a
252 temperature and/or angular velocity gradient in response to the
253 applied flux. By fitting the radial temperature gradient, it is
254 straightforward to obtain the bulk thermal conductivity,
255 \begin{equation}
256 J_r = -\lambda \left( \frac{\partial T}{\partial r}\right)
257 \end{equation}
258 from the radial kinetic energy flux $(J_r)$ that was applied during
259 the simulation. In practice, it is significantly easier to use the
260 integrated form of Fourier's law for spherical geometries. In
261 sections \ref{sec:thermal} -- \ref{sec:rotation} we outline ways of
262 obtaining interfacial transport coefficients from these RNEMD
263 simulations.
264
265 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
266 % **COMPUTATIONAL DETAILS**
267 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268 \section{Computational Details}
269
270 The new VSS-RNEMD methodology for non-periodic system geometries has
271 been implemented in our group molecular dynamics code,
272 OpenMD.\cite{Meineke:2005gd,openmd} We have tested the new method to
273 calculate the thermal conductance of a gold nanoparticle and SPC/E
274 water cluster, and compared the results with previous bulk RNEMD
275 values, as well as experiment. We have also investigated the
276 interfacial thermal conductance and interfacial rotational friction
277 for gold nanostructures solvated in hexane as a function of
278 nanoparticle size and shape.
279
280 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
281 % FORCE FIELD PARAMETERS
282 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
283 \subsection{Force field}
284
285 Gold -- gold interactions are described by the quantum Sutton-Chen
286 (QSC) model.\cite{PhysRevB.59.3527} The QSC parameters are tuned to
287 experimental properties such as density, cohesive energy, and elastic
288 moduli and include zero-point quantum corrections.
289
290 The SPC/E water model~\cite{Berendsen87} is particularly useful for
291 validation of conductivities and shear viscosities. This model has
292 been used to previously test other RNEMD and NEMD approaches, and
293 there are reported values for thermal conductivies and shear
294 viscosities at a wide range of thermodynamic conditions that are
295 available for direct comparison.\cite{Bedrov:2000,Kuang:2010if}
296
297 Hexane molecules are described by the TraPPE united atom
298 model,\cite{TraPPE-UA.alkanes} which provides good computational
299 efficiency and reasonable accuracy for bulk thermal conductivity
300 values. In this model, sites are located at the carbon centers for
301 alkyl groups. Bonding interactions, including bond stretches and bends
302 and torsions, were used for intra-molecular sites closer than 3
303 bonds. For non-bonded interactions, Lennard-Jones potentials were
304 used. We have previously utilized both united atom (UA) and all-atom
305 (AA) force fields for thermal
306 conductivity,\cite{Kuang:2011ef,Kuang:2012fe,Stocker:2013cl} and since
307 the united atom force fields cannot populate the high-frequency modes
308 that are present in AA force fields, they appear to work better for
309 modeling thermal conductance at metal/ligand interfaces.
310
311 Gold -- hexane nonbonded interactions are governed by pairwise
312 Lennard-Jones parameters derived from Vlugt \emph{et
313 al}.\cite{vlugt:cpc2007154} They fitted parameters for the
314 interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the
315 effective potential of Hautman and Klein for the Au(111)
316 surface.\cite{hautman:4994}
317
318 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
319 % NON-PERIODIC DYNAMICS
320 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
321 % \subsection{Dynamics for non-periodic systems}
322 %
323 % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
324 % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
325 % compressibilities, which are typically problematic for traditional affine transform methods. We have had
326 % success applying this method to several different systems including bare metal nanoparticles, liquid water
327 % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
328 % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
329 % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
330 % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
331 % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
332 % subject to standard Newtonian dynamics.
333
334 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335 % SIMULATION PROTOCOL
336 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
337 \subsection{Simulation protocol}
338
339 In all cases, systems were equilibrated under non-periodic
340 isobaric-isothermal (NPT) conditions -- using the Langevin Hull
341 methodology\cite{Vardeman2011} -- before any non-equilibrium methods
342 were introduced. For heterogeneous systems, the gold nanoparticles and
343 ellipsoids were created from a bulk fcc lattice and were thermally
344 equilibrated before being solvated in hexane. Packmol\cite{packmol}
345 was used to solvate previously equilibrated gold nanostructures within
346 a spherical droplet of hexane.
347
348 Once equilibrated, thermal or angular momentum fluxes were applied for
349 1 - 2 ns, until stable temperature or angular velocity gradients had
350 developed. Systems containing liquids were run under moderate pressure
351 (5 atm) and temperatures (230 K) to avoid the formation of a vapor
352 layer at the boundary of the cluster. Pressure was applied to the
353 system via the non-periodic Langevin Hull.\cite{Vardeman2011} However,
354 thermal coupling to the external temperature and pressure bath was
355 removed to avoid interference with the imposed RNEMD flux.
356
357 Because the method conserves \emph{total} angular momentum, systems
358 which contain a metal nanoparticle embedded in a significant volume of
359 solvent will still experience nanoparticle diffusion inside the
360 solvent droplet. To aid in computing the rotational friction in these
361 systems, a single gold atom at the origin of the coordinate system was
362 assigned a mass $10,000 \times$ its original mass. The bonded and
363 nonbonded interactions for this atom remain unchanged and the heavy
364 atom is excluded from the RNEMD exchanges. The only effect of this
365 gold atom is to effectively pin the nanoparticle at the origin of the
366 coordinate system, while still allowing for rotation. For rotation of
367 the gold ellipsoids we added two of these heavy atoms along the axis
368 of rotation, separated by an equal distance from the origin of the
369 coordinate system. These heavy atoms prevent off-axis tumbling of the
370 nanoparticle and allow for measurement of rotational friction relative
371 to a particular axis of the ellipsoid.
372
373 Angular velocity data was collected for the heterogeneous systems
374 after a brief period of imposed flux to initialize rotation of the
375 solvated nanostructure. Doing so ensures that we overcome the initial
376 static friction and calculate only the \emph{dynamic} interfacial
377 rotational friction.
378
379 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
380 % THERMAL CONDUCTIVITIES
381 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
382 \subsection{Thermal conductivities}
383 \label{sec:thermal}
384
385 To compute the thermal conductivities of bulk materials, the
386 integrated form of Fourier's Law of heat conduction in radial
387 coordinates can be used to obtain an expression for the heat transfer
388 rate between concentric spherical shells:
389 \begin{equation}
390 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
391 \label{eq:Q}
392 \end{equation}
393 The heat transfer rate, $q_r$, is constant in spherical geometries,
394 while the heat flux, $J_r$ depends on the surface area of the two
395 shells. $\lambda$ is the thermal conductivity, and $T_{a,b}$ and
396 $r_{a,b}$ are the temperatures and radii of the two concentric RNEMD
397 regions, respectively.
398
399 A kinetic energy flux is created using VSS-RNEMD moves, and the
400 temperature in each of the radial shells is recorded. The resulting
401 temperature profiles are analyzed to yield information about the
402 interfacial thermal conductance. As the simulation progresses, the
403 VSS moves are performed on a regular basis, and the system develops a
404 thermal or velocity gradient in response to the applied flux. Once a
405 stable thermal gradient has been established between the two regions,
406 the thermal conductivity, $\lambda$, can be calculated using a linear
407 regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
408
409 \begin{equation}
410 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
411 \label{eq:lambda}
412 \end{equation}
413
414 The rate of heat transfer, $q_r$, between the two RNEMD regions is
415 easily obtained from either the applied kinetic energy flux and the
416 area of the smaller of the two regions, or from the total amount of
417 transferred kinetic energy and the run time of the simulation.
418 \begin{equation}
419 q_r = J_r A = \frac{KE}{t}
420 \label{eq:heat}
421 \end{equation}
422
423 \subsubsection{Thermal conductivity of nanocrystalline gold}
424 Calculated values for the thermal conductivity of a 40 \AA$~$ radius
425 gold nanoparticle (15707 atoms) at a range of kinetic energy flux
426 values are shown in Table \ref{table:goldTC}. For these calculations,
427 the hot and cold slabs were excluded from the linear regression of the
428 thermal gradient.
429
430 \begin{table}
431 \caption{Calculated thermal gradients of a crystalline gold
432 nanoparticle of radius 40 \AA\ subject to a range of applied
433 kinetic energy fluxes. Calculations were performed at an average
434 temperature of 300 K. Gold-gold interactions are described by the
435 Quantum Sutton-Chen potential.}
436 \label{table:goldTC}
437 \begin{tabular}{cc}
438 \\ \hline \hline
439 {$J_r$} & {$\langle dT/dr \rangle$} \\
440 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K /
441 \AA)} \\ \hline \\
442 3.25$\times 10^{-6}$ & 0.114 \\
443 6.50$\times 10^{-6}$ & 0.232 \\
444 1.30$\times 10^{-5}$ & 0.449 \\
445 3.25$\times 10^{-5}$ & 1.180 \\
446 6.50$\times 10^{-5}$ & 2.339 \\ \hline \hline
447 \end{tabular}
448 \end{table}
449
450 The measured thermal gradients $\langle dT/dr \rangle$ are linearly
451 dependent on the applied kinetic energy flux $J_r$. The calculated
452 thermal conductivity value, $\lambda = 1.004 \pm
453 0.009$~W~/~m~$\cdot$~K compares well with previous {\it bulk} QSC
454 values of 1.08~--~1.26~W~/~m~$\cdot$~K\cite{Kuang:2010if}, though
455 still significantly lower than the experimental value of
456 320~W~/~m~$\cdot$~K, as the QSC force field neglects the significant
457 electronic contributions to thermal conductivity. We note that there
458 is only a minimal effect on the computed thermal conductivity of gold
459 when comparing periodic (bulk) simulations carried out at the same
460 densities as those used here.
461
462 \subsubsection{Thermal conductivity of a droplet of SPC/E water}
463
464 Calculated values for the thermal conductivity of a cluster of 6912
465 SPC/E water molecules were computed with a range of applied kinetic
466 energy fluxes. As with the gold nanoparticle, the measured slopes were
467 linearly dependent on the applied kinetic energy flux $J_r$, and the
468 RNEMD regions were excluded from the $\langle dT/dr \rangle$ fit (see
469 Fig. \ref{fig:lambda_G}).
470
471 The computed mean value of the thermal conductivity, $\lambda = 0.884
472 \pm 0.013$~W~/~m~$\cdot$~K, compares well with previous
473 non-equilibrium molecular dynamics simulations of bulk SPC/E that were
474 carried out in periodic geometries.\cite{Zhang2005,Romer2012} These
475 simualtions gave conductivity values from 0.81--0.87~W~/~m~$\cdot$~K,
476 and all of the simulated values are approximately 30-45\% higher than the
477 experimental thermal conductivity of water, which has been measured at
478 0.61~W~/~m~$\cdot$~K.\cite{WagnerKruse}
479
480 \begin{figure}
481 \includegraphics[width=\linewidth]{figures/lambda_G}
482 \caption{Upper panel: temperature profile of a typical SPC/E water
483 droplet. The RNEMD simulation transfers kinetic energy from region
484 A to region B, and the system responds by creating a temperature
485 gradient (circles). The fit to determine $\langle dT/dr \rangle$ is
486 carried out between the two RNEMD exchange regions. Lower panel:
487 temperature profile for a solvated Au nanoparticle ($r = 20$\AA).
488 The temperature differences used to compute the total Kapitza
489 resistance (and interfacial thermal conductance) are indicated with
490 arrows.}
491 \label{fig:lambda_G}
492 \end{figure}
493
494
495 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
496 % INTERFACIAL THERMAL CONDUCTANCE
497 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
498 \subsection{Interfacial thermal conductance}
499 \label{sec:interfacial}
500
501 The interfacial thermal conductance, $G$, of a heterogeneous interface
502 located at $r_0$ can be understood as the change in thermal
503 conductivity in a direction normal $(\mathbf{\hat{n}})$ to the
504 interface,
505 \begin{equation}
506 G = \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{r_0}
507 \end{equation}
508 For heterogeneous systems such as the solvated nanoparticle shown in
509 Figure \ref{fig:NP20}, the interfacial thermal conductance at the
510 surface of the nanoparticle can be determined using a kinetic energy
511 flux applied using the RNEMD method developed above.
512
513 In spherical geometries, it is most convenient to begin by finding the
514 Kapitza or interfacial thermal resistance for a thin spherical shell,
515 \begin{equation}
516 R_K = \frac{1}{G} = \frac{\Delta
517 T}{J_r}
518 \end{equation}
519 where $\Delta T$ is the temperature drop from the interior to the
520 exterior of the shell. For two concentric shells, the kinetic energy
521 flux ($J_r$) is not the same (as the surface areas are not the same),
522 but the heat transfer rate, $q_r = J_r A$ is constant. The thermal
523 resistance of a shell with interior radius $r$ is most conveniently
524 written as
525 \begin{equation}
526 R_K = \frac{1}{q_r} \Delta T 4 \pi r^2.
527 \label{eq:RK}
528 \end{equation}
529
530 \begin{figure}
531 \includegraphics[width=\linewidth]{figures/NP20}
532 \caption{A gold nanoparticle with a radius of 20 \AA$\,$ solvated in
533 TraPPE-UA hexane. A kinetic energy flux is applied between the
534 nanoparticle and an outer shell of solvent to obtain the interfacial
535 thermal conductance, $G$, and the interfacial rotational resistance
536 $\Xi^{rr}$ is determined using an angular momentum flux. }
537 \label{fig:NP20}
538 \end{figure}
539
540 To model the thermal conductance across a wide interface (or multiple
541 interfaces) it is useful to consider concentric shells as resistors
542 wired in series. The resistance of the shells is then additive, and
543 the interfacial thermal conductance is the inverse of the total
544 Kapitza resistance:
545 \begin{equation}
546 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
547 T_i\right) 4 \pi r_i^2
548 \end{equation}
549 This series can be extended for any number of adjacent shells,
550 allowing for the calculation of the interfacial thermal conductance
551 for interfaces of considerable thickness, such as self-assembled
552 ligand monolayers on a metal surface.
553
554 \subsubsection{Interfacial thermal conductance of solvated gold
555 nanoparticles}
556 Calculated interfacial thermal conductance ($G$) values for three
557 sizes of gold nanoparticles and a flat Au(111) surface solvated in
558 TraPPE-UA hexane~\cite{Stocker:2013cl} are shown in Table
559 \ref{table:G}.
560
561 \begin{table}
562 \caption{Calculated interfacial thermal conductance ($G$) values for
563 gold nanoparticles of varying radii solvated in TraPPE-UA
564 hexane. The nanoparticle $G$ values are compared to previous
565 simulation results for a Au(111) interface in TraPPE-UA hexane.}
566 \label{table:G}
567 \begin{tabular}{cc}
568 \hline \hline
569 {Nanoparticle Radius} & {$G$}\\
570 {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
571 20 & {47.1 $\pm$ 5.0} \\
572 30 & {45.4 $\pm$ 1.2} \\
573 40 & {46.5 $\pm$ 2.1} \\
574 \hline
575 Au(111) & {30.2} \\
576 \hline \hline
577 \end{tabular}
578 \end{table}
579
580 The introduction of surface curvature increases the interfacial
581 thermal conductance by a factor of approximately $1.5$ relative to the
582 flat interface, although there is no apparent size-dependence in the
583 $G$ values obtained from our calculations. This is quite different
584 behavior than the size-dependence observed by Lervik \textit{et
585 al.}\cite{Lervik:2009fk} in their NEMD simulations of decane
586 droplets in water. In the liquid/liquid case, the surface tension is
587 strongly dependent on the droplet size, and the interfacial
588 resistivity increases with surface tension. In our case, the surface
589 tension of the solid / liquid interface does not have a strong
590 dependence on the particle radius at these particle sizes.
591
592 Simulations of larger nanoparticles may yet demonstrate limiting $G$
593 values close to the flat Au(111) slab, although any spherical particle
594 will have a significant fraction of the surface atoms in non-111
595 facets. It is not clear at this point whether the increase in thermal
596 conductance for the spherical particles is due to the increased
597 population of undercoordinated surface atoms, or to increased solid
598 angle space available for phonon scattering into the solvent. These
599 two possibilities do open up some interesting avenues for
600 investigation.
601
602 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
603 % INTERFACIAL ROTATIONAL FRICTION
604 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
605 \subsection{Interfacial rotational friction}
606 \label{sec:rotation}
607 The interfacial rotational resistance tensor, $\Xi^{rr}$, can be
608 calculated for heterogeneous nanostructure/solvent systems by applying
609 an angular momentum flux between the solvated nanostructure and a
610 spherical shell of solvent at the outer edge of the cluster. An
611 angular velocity gradient develops in response to the applied flux,
612 causing the nanostructure and solvent shell to rotate in opposite
613 directions about a given axis.
614
615 \begin{figure}
616 \includegraphics[width=\linewidth]{figures/E25-75}
617 \caption{A gold prolate ellipsoid of length 65 \AA$\,$ and width 25
618 \AA$\,$ solvated by TraPPE-UA hexane. An angular momentum flux is
619 applied between the ellipsoid and an outer shell of solvent.}
620 \label{fig:E25-75}
621 \end{figure}
622
623 Analytical solutions for the diagonal elements of the rotational
624 resistance tensor for solvated spherical body of radius $r$ under
625 ideal stick boundary conditions can be estimated using Stokes' law
626 \begin{equation}
627 \Xi^{rr}_{stick} = 8 \pi \eta r^3,
628 \label{eq:Xisphere}
629 \end{equation}
630 where $\eta$ is the dynamic viscosity of the surrounding solvent.
631
632 For general ellipsoids with semiaxes $\alpha$, $\beta$, and $\gamma$,
633 Perrin's extension of Stokes' law provides exact solutions for
634 symmetric prolate $(\alpha \geq \beta = \gamma)$ and oblate $(\alpha <
635 \beta = \gamma)$ ellipsoids under ideal stick conditions. The Perrin
636 elliptic integral parameter,
637 \begin{equation}
638 S = \frac{2}{\sqrt{\alpha^2 - \beta^2}} \ln \left[ \frac{\alpha + \sqrt{\alpha^2 - \beta^2}}{\beta} \right].
639 \label{eq:S}
640 \end{equation}
641
642 For a prolate ellipsoidal nanoparticle (see Fig. \ref{fig:E25-75}),
643 the rotational resistance tensor $\Xi^{rr}$ is a $3 \times 3$ diagonal
644 matrix with elements
645 \begin{align}
646 \Xi^{rr}_\alpha =& \frac{32 \pi}{3} \eta \frac{ \left(
647 \alpha^2 - \beta^2 \right) \beta^2}{2\alpha - \beta^2 S}
648 \\ \nonumber \\
649 \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},
650 \label{eq:Xirr}
651 \end{align}
652 corresponding to rotation about the long axis ($\alpha$), and each of
653 the equivalent short axes ($\beta$ and $\gamma$), respectively.
654
655 Previous VSS-RNEMD simulations of the interfacial friction of the
656 planar Au(111) / hexane interface have shown that the flat interface
657 exists within slip boundary conditions.\cite{Kuang:2012fe} Hu and
658 Zwanzig\cite{Zwanzig} investigated the rotational friction
659 coefficients for spheroids under slip boundary conditions and obtained
660 numerial results for a scaling factor to be applied to
661 $\Xi^{rr}_{\mathit{stick}}$ as a function of ratio of the shorter
662 semiaxes to the longer. Under slip conditions, rotation of any
663 sphere, and rotation of a prolate ellipsoid about its long axis does
664 not displace any solvent, so the resulting $\Xi^{rr}_{\mathit{slip}}$
665 approaches $0$. For rotation of a prolate ellipsoid (with the aspect
666 ratio shown here) about its short axis, Hu and Zwanzig obtained
667 $\Xi^{rr}_{\mathit{slip}}$ values of $35.9\%$ of the analytical
668 $\Xi^{rr}_{\mathit{stick}}$ result. This reduced rotational
669 resistance accounts for the reduced interfacial friction under slip
670 boundary conditions.
671
672 The measured rotational friction coefficient,
673 $\Xi^{rr}_{\mathit{meas}}$ at the interface can be extracted from
674 non-periodic VSS-RNEMD simulations quite easily using the total
675 applied torque ($\tau$) and the observed angular velocity of the solid
676 structure ($\omega_\mathrm{solid}$),
677 \begin{equation}
678 \Xi^{rr}_{\mathit{meas}} = \frac{\tau}{\omega_\mathrm{solid}}
679 \label{eq:Xieff}
680 \end{equation}
681
682 The total applied torque required to overcome the interfacial friction
683 and maintain constant rotation of the gold is
684 \begin{equation}
685 \tau = \frac{j_r(\mathbf{L}) \cdot A}{2}
686 \label{eq:tau}
687 \end{equation}
688 where $j_r(\mathbf{L})$ is the applied angular momentum flux, $A$ is
689 the surface area of the solid nanoparticle, and the factor of $2$ is
690 present because the torque is exerted on both the particle and the
691 counter-rotating solvent shell.
692
693 \subsubsection{Rotational friction for gold nanostructures in hexane}
694
695 Table \ref{table:couple} shows the calculated rotational friction
696 coefficients $\Xi^{rr}$ for spherical gold nanoparticles and a prolate
697 ellipsoidal gold nanorod solvated in TraPPE-UA hexane. An angular
698 momentum flux was applied between an outer shell of solvent that was
699 at least 20 \AA\ away from the surface of the particle (region $A$)
700 and the gold nanostructure (region $B$). Dynamic viscosity estimates
701 $(\eta)$ for TraPPE-UA hexane under these particular temperature and
702 pressure conditions was determined by applying a traditional VSS-RNEMD
703 linear momentum flux to a periodic box of this solvent.
704
705 \begin{table}
706 \caption{Comparison of rotational friction coefficients under ideal stick
707 ($\Xi^{rr}_{\mathit{stick}}$) and
708 slip ($\Xi^{rr}_{\mathit{slip}}$) conditions, along with the measured
709 ($\Xi^{rr}_{\mathit{meas}}$) rotational friction coefficients of gold
710 nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid
711 is oriented with the long axis along the $z$ direction.}
712 \label{table:couple}
713 \begin{tabular}{lccccc}
714 \\ \hline \hline
715 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{meas}}$} & {$\Xi^{rr}_{\mathit{meas}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
716 {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
717 Sphere (r = 20 \AA) & {$x = y = z$} & {3314} & 0 & {2386 $\pm$ 14} & {0.720 $\pm$ 0.004}\\
718 Sphere (r = 30 \AA) & {$x = y = z$} & {11749}& 0 & {8415 $\pm$ 274} & {0.716 $\pm$ 0.023}\\
719 Sphere (r = 40 \AA) & {$x = y = z$} & {34464}& 0 & {47544 $\pm$ 3051} & {1.380 $\pm$ 0.089}\\
720 Prolate Ellipsoid & {$x = y$} & {4991} & {1792} & {3128 $\pm$ 166} & {0.627 $\pm$ 0.033}\\
721 Prolate Ellipsoid & {$z$} & {1993} & 0 & {1590 $\pm$ 30} & {0.798 $\pm$ 0.015}
722 \\ \hline \hline
723 \end{tabular}
724 \end{table}
725
726 The results for $\Xi^{rr}_{\mathit{meas}}$ show that, in contrast with
727 the flat Au(111) / hexane interface, gold nanostructures solvated by
728 hexane are closer to stick than slip boundary conditions. At these
729 length scales, the nanostructures are not perfect spheroids due to
730 atomic `roughening' of the surface and therefore experience increased
731 interfacial friction, deviating significantly from the ideal slip
732 case. The 20 and 30~\AA\ radius nanoparticles experience approximately
733 70\% of the ideal stick boundary interfacial friction. Rotation of the
734 ellipsoid about its long axis more closely approaches the stick limit
735 than rotation about the short axis, which at first seems
736 counterintuitive. However, the `propellor' motion caused by rotation
737 around the short axis may exclude solvent from the rotation cavity or
738 move a sufficient amount of solvent along with the gold that a smaller
739 interfacial friction is actually experienced.
740
741 The largest nanoparticle
742 (40 \AA$\,$ radius) appears to experience interfacial friction in
743 excess of ideal stick conditions. Although the solvent velocity
744 field does not appear in the calculation of $\Xi^{rr}$, we do note
745 that the smaller particles have solvent velocity fields that can be
746 fit with the $v(r,\theta) = \left( A r + B/r^2
747 \right) \sin\theta $ form,\cite{Jeffery:1915fk} while the solvent velocity fields
748 for the 40 \AA\ radius particle approaches the infinite fluid
749 solution and can be fit well with only the second term.
750
751 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
752 % **DISCUSSION**
753 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
754 \section{Discussion}
755
756 We have adapted the VSS-RNEMD methodology so that it can be used to
757 apply kinetic energy and angular momentum fluxes in explicitly
758 non-periodic simulation geometries. Non-periodic VSS-RNEMD preserves
759 the strengths of the original periodic variant, specifically Boltzmann
760 velocity distributions and minimal thermal anisotropy, while extending
761 the constraints to conserve total energy and total \emph{angular}
762 momentum. This method also maintains the ability to impose the kinetic
763 energy and angular momentum fluxes either jointly or individually.
764
765 This method enables steady-state calculation of interfacial thermal
766 conductance and interfacial rotational friction in heterogeneous
767 clusters. We have demonstrated the abilities of the method on some
768 familiar systems (nanocrystalline gold, SPC/E water) as well as on
769 interfaces that have been studied only in planar geometries. For the
770 bulk systems that are well-understood in periodic geometries, the
771 nonperiodic VSS-RNEMD provides very similar estimates of the thermal
772 conductivity to other simulation methods.
773
774 For nanoparticle-to-solvent heat transfer, we have observed that the
775 interfacial conductance exhibits a $1.5\times$ increase over the
776 planar Au(111) interface, although we do not see any dependence of the
777 conductance on the particle size. Because the surface tension effects
778 present in liquid/liquid heat transfer are not strong contributors
779 here, we suggest two possible mechanisms for the $1.5\times$ increase
780 over the planar surface: (1) the nanoparticles have an increased
781 population of undercoordinated surface atoms that are more efficient
782 at transferring vibrational energy to the solvent, or (2) there is
783 increased solid angle space available for phonon scattering into the
784 solvent. The second of these explanations would have a significant
785 dependence on the radius of spherical particles, although crystalline
786 faceting of the particles may be enough to reduce this dependence on
787 particle radius. These two possibilities do open up some interesting
788 avenues for further exploration.
789
790 One area that this method opens up that was not available for periodic
791 simulation cells is direct computation of the solid/liquid rotational
792 friction coefficients of nanostructures. The systems we have
793 investigated (gold nanospheres and prolate ellipsoids) have analytic
794 stick and approximate slip solutions provided via hydrodynamics, and
795 our molecular simulations indicate that the bare metallic
796 nanoparticles are closer to the stick than they are to slip
797 conditions. This is markedly different behavior than we see for the
798 solid / liquid friction at planar Au(111) interfaces, where the
799 solvents experience a nearly-laminar flow over the surface in previous
800 simulations. Schmidt and Skinner previously computed the behavior of
801 spherical tag particles in molecular dynamics simulations, and showed
802 that {\it slip} boundary conditions for translational resistance may
803 be more appropriate for molecule-sized spheres embedded in a sea of
804 spherical solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx} The
805 size of the gold nanoparticles studied here and the structuring of the
806 surfaces of the particles appears to bring the behavior closer to
807 stick boundaries. Exactly where the slip-stick crossover takes place
808 is also an interesting avenue for exploration.
809
810 The ability to interrogate explicitly non-periodic effects
811 (e.g. surface curvature, edge effects between facets, and the splay of
812 ligand surface groups) on interfacial transport means that this method
813 can be applied to systems of broad experimental and theoretical
814 interest.
815
816 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
817 % **ACKNOWLEDGMENTS**
818 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
819 \begin{acknowledgement}
820 The authors thank Dr. Shenyu Kuang for helpful discussions. Support
821 for this project was provided by the National Science Foundation
822 under grant CHE-0848243. Computational time was provided by the
823 Center for Research Computing (CRC) at the University of Notre Dame.
824 \end{acknowledgement}
825
826
827 \newpage
828
829 \bibliography{nonperiodicVSS}
830
831 %\end{doublespace}
832 \end{document}

Properties

Name Value
svn:executable