ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4064
Committed: Thu Mar 13 20:44:45 2014 UTC (10 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 36932 byte(s)
Log Message:
edits

File Contents

# User Rev Content
1 kstocke1 3927 \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 gezelter 4064 \usepackage{wrapfig}
20 kstocke1 3927 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
21     \usepackage{url}
22    
23 gezelter 4063 \title{A method for creating thermal and angular momentum fluxes in
24     non-periodic simulations}
25 kstocke1 3927
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 gezelter 4063 \begin{tocentry}
34 gezelter 4064 \includegraphics[width=3.6cm]{figures/NP20}
35 gezelter 4063 \end{tocentry}
36    
37    
38 kstocke1 3927 \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 gezelter 4063 %\date{\today}
50 kstocke1 3927
51 gezelter 4063 %\maketitle
52 kstocke1 3927
53 gezelter 4063 %\begin{doublespace}
54 kstocke1 3927
55     \begin{abstract}
56    
57 gezelter 4063 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 kstocke1 3927
74     \end{abstract}
75    
76     \newpage
77    
78     %\narrowtext
79    
80     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
81     % **INTRODUCTION**
82     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83     \section{Introduction}
84    
85 gezelter 4063 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 kstocke1 3927
105 gezelter 4063 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 kstocke1 4009
125 gezelter 4060 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 gezelter 4063 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 kstocke1 4009
139 gezelter 4063 To date, however, the RNEMD methods have only been used in periodic
140 gezelter 4060 simulation cells where the exchange regions are physically separated
141 gezelter 4063 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 gezelter 4060 non-periodic simulations, and use the method to compute some thermal
150     transport and solid-liquid friction at the surfaces of spherical and
151 gezelter 4063 ellipsoidal nanoparticles.
152 gezelter 4060
153 kstocke1 4009 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
154     % **METHODOLOGY**
155     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156 gezelter 4063 \section{Velocity shearing and scaling (VSS) for non-periodic systems}
157 kstocke1 4009
158 gezelter 4063 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 kstocke1 4009
166 kstocke1 3994 \begin{figure}
167     \includegraphics[width=\linewidth]{figures/npVSS}
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 gezelter 3977
176 gezelter 4063 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 gezelter 3977
188 gezelter 4063 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 kstocke1 4003
200 gezelter 4063 Once the values of the scaling and shearing factors are known, the
201     velocity changes are applied,
202 gezelter 3977 \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 kstocke1 4009 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 gezelter 3977 $\mathbf{j}_r(\mathbf{L})$,
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 gezelter 4063 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia
226     tensor for each of the two shells.
227 gezelter 3977
228 gezelter 4063 To simultaneously impose a thermal flux ($J_r$) between the shells we
229     use energy conservation constraints,
230 gezelter 3977 \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 kstocke1 4009 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 gezelter 3977
246 gezelter 4063 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 gezelter 4064 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 gezelter 3977
261 kstocke1 4003 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262     % **COMPUTATIONAL DETAILS**
263     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264     \section{Computational Details}
265    
266 gezelter 4063 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 kstocke1 4009
276 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277 kstocke1 4009 % FORCE FIELD PARAMETERS
278 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
279 gezelter 4063 \subsection{Force field}
280 kstocke1 3927
281 gezelter 4063 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 kstocke1 3927
286 gezelter 4063 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 kstocke1 3947
293 gezelter 4063 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 kstocke1 3947
307 gezelter 4063 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 kstocke1 4009
314 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
315 kstocke1 4009 % NON-PERIODIC DYNAMICS
316 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
317 kstocke1 4009 % \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 kstocke1 3947
330 kstocke1 4009 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
331     % SIMULATION PROTOCOL
332     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
333     \subsection{Simulation protocol}
334 kstocke1 3947
335 gezelter 4063 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 kstocke1 3947
344 gezelter 4063 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 kstocke1 3947
353 gezelter 4063 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 kstocke1 3947
369 gezelter 4063 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 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
376     % THERMAL CONDUCTIVITIES
377     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378     \subsection{Thermal conductivities}
379 gezelter 4064 \label{sec:thermal}
380 kstocke1 3947
381 gezelter 4064 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 kstocke1 3947 \begin{equation}
386 gezelter 4064 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
387 kstocke1 4003 \label{eq:Q}
388 kstocke1 3947 \end{equation}
389 gezelter 4064 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 kstocke1 3947
395 gezelter 4064 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 gezelter 4063 regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
404 kstocke1 3991
405     \begin{equation}
406 gezelter 4064 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
407 kstocke1 4003 \label{eq:lambda}
408 kstocke1 3991 \end{equation}
409    
410 gezelter 4064 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 kstocke1 3991 \begin{equation}
415 gezelter 4064 q_r = J_r A = \frac{KE}{t}
416 kstocke1 4003 \label{eq:heat}
417 kstocke1 3991 \end{equation}
418    
419 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
420     % INTERFACIAL THERMAL CONDUCTANCE
421     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
422     \subsection{Interfacial thermal conductance}
423 gezelter 4064 \label{sec:interfacial}
424 kstocke1 3947
425 gezelter 4064 The interfacial thermal conductance, $G$, of a heterogeneous interface
426     located at $r_0$ can be understood as the change in thermal
427     conductivity in a direction normal $(\mathbf{\hat{n}})$ to the
428     interface,
429     \begin{equation}
430     G = \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{r_0}
431     \end{equation}
432     For heterogeneous systems such as the solvated nanoparticle shown in
433     Figure \ref{fig:NP20}, the interfacial thermal conductance at the
434     surface of the nanoparticle can be determined using a kinetic energy
435     flux applied using the RNEMD method developed above. It is most
436     convenient to compute the Kaptiza or interfacial thermal resistance
437     for a thin spherical shell,
438     \begin{equation}
439     R_K = \frac{1}{G} = \frac{\Delta
440     T}{J_r}
441     \end{equation}
442     where $\Delta T$ is the temperature drop from the interior to the
443     exterior of the shell. For two neighboring shells, the kinetic energy
444     flux ($J_r$) is not the same (as the surface areas are not the same),
445     but the heat transfer rate, $q_r = J_r A$ is constant. The thermal
446     resistance of a shell with interior radius $r$ is most conveniently
447     written as
448     \begin{equation}
449     R_K = \frac{1}{q_r} \Delta T 4 \pi r^2.
450     \label{eq:RK}
451     \end{equation}
452    
453    
454 kstocke1 4009 \begin{figure}
455     \includegraphics[width=\linewidth]{figures/NP20}
456 gezelter 4064 \caption{A gold nanoparticle with a radius of 20 \AA$\,$ solvated in
457     TraPPE-UA hexane. A kinetic energy flux is applied between the
458     nanoparticle and an outer shell of solvent to obtain the interfacial
459     thermal conductance, $G$, and the interfacial rotational resistance
460     $\Xi^{rr}$ is determined using an angular momentum flux. }
461 kstocke1 4009 \label{fig:NP20}
462     \end{figure}
463 kstocke1 4003
464 gezelter 4064 To model the thermal conductance across an interface (or multiple
465     interfaces) it is useful to consider the shells as resistors wired in
466     series. The resistance of the shells is then additive, and the
467     interfacial thermal conductance is the inverse of the total Kapitza
468     resistance:
469 kstocke1 3947 \begin{equation}
470 gezelter 4064 \frac{1}{G} = R_\mathrm{total} = \frac{1}{q_r} \sum_i \left(T_{i+i} -
471     T_i\right) 4 \pi r_i^2
472 kstocke1 3947 \end{equation}
473 gezelter 4064 This series can be expanded for any number of adjacent shells,
474     allowing for the calculation of the interfacial thermal conductance
475     for interfaces of considerable thickness, such as self-assembled
476     ligand monolayers on a metal surface.
477 kstocke1 3947
478 kstocke1 4003
479 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
480 kstocke1 4009 % INTERFACIAL ROTATIONAL FRICTION
481 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
482 kstocke1 4009 \subsection{Interfacial rotational friction}
483 gezelter 4064 \label{sec:rotation}
484     The interfacial rotational resistance tensor, $\Xi^{rr}$, can be
485     calculated for heterogeneous nanostructure/solvent systems by applying
486     an angular momentum flux between the solvated nanostructure and a
487     spherical shell of solvent at the outer edge of the cluster. An
488     angular velocity gradient develops in response to the applied flux,
489     causing the nanostructure and solvent shell to rotate in opposite
490     directions about a given axis.
491 kstocke1 3947
492 kstocke1 4009 \begin{figure}
493     \includegraphics[width=\linewidth]{figures/E25-75}
494 gezelter 4064 \caption{A gold prolate ellipsoid of length 65 \AA$\,$ and width 25
495     \AA$\,$ solvated by TraPPE-UA hexane. An angular momentum flux is
496     applied between the ellipsoid and an outer shell of solvent.}
497 kstocke1 4009 \label{fig:E25-75}
498     \end{figure}
499    
500 gezelter 4064 Analytical solutions for the diagonal elements of the rotational
501     resistance tensor for solvated spherical body of radius $r$ under
502     ideal stick boundary conditions can be estimated using Stokes' law
503 kstocke1 3947 \begin{equation}
504 gezelter 4064 \Xi^{rr}_{stick} = 8 \pi \eta r^3,
505     \label{eq:Xisphere}
506 kstocke1 3947 \end{equation}
507 gezelter 4064 where $\eta$ is the dynamic viscosity of the surrounding solvent.
508 kstocke1 3947
509 gezelter 4064 For general ellipsoids with semiaxes $\alpha$, $\beta$, and $\gamma$,
510     Perrin's extension of Stokes' law provides exact solutions for
511     symmetric prolate $(\alpha \geq \beta = \gamma)$ and oblate $(\alpha <
512     \beta = \gamma)$ ellipsoids under ideal stick conditions. For
513     simplicity, we define the Perrin Factor, $S$,
514 kstocke1 3947
515     \begin{equation}
516 gezelter 4064 S = \frac{2}{\sqrt{\alpha^2 - \beta^2}} \ln \left[ \frac{\alpha + \sqrt{\alpha^2 - \beta^2}}{\beta} \right].
517 kstocke1 4003 \label{eq:S}
518 kstocke1 3947 \end{equation}
519    
520 gezelter 4064 For a prolate ellipsoidal nanoparticle (see Fig. \ref{fig:E25-75}),
521     the rotational resistance tensor $\Xi^{rr}$ is a $3 \times 3$ diagonal
522     matrix with elements
523 kstocke1 3947 \begin{equation}
524 gezelter 4064 \Xi^{rr}_\alpha = \frac{32 \pi}{3} \eta \frac{ \left( \alpha^2 - \beta^2 \right) \beta^2}{2\alpha - \beta^2 S}
525 kstocke1 4003 \label{eq:Xia}
526 kstocke1 3991 \end{equation}\vspace{-0.45in}\\
527     \begin{equation}
528 gezelter 4064 \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},
529 kstocke1 4003 \label{eq:Xibc}
530 kstocke1 3947 \end{equation}
531 gezelter 4064 corresponding to rotation about the long axis ($\alpha$), and each of
532     the equivalent short axes ($\beta$ and $\gamma$), respectively.
533 kstocke1 3991
534 gezelter 4064 Previous VSS-RNEMD simulations of the interfacial friction of the
535     planar Au(111) / hexane interface have shown that the interface exists
536     within slip boundary conditions.\cite{Kuang:2012fe} Hu and
537     Zwanzig\cite{Zwanzig} investigated the rotational friction
538     coefficients for spheroids under slip boundary conditions and obtained
539     numerial results for a scaling factor to be applied to
540     $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau$, the ratio of the
541     shorter semiaxes and the longer semiaxis of the spheroid. For the
542     sphere and prolate ellipsoid shown here, the values of $\tau$ are $1$
543     and $0.3939$, respectively. Under ``slip'' conditions,
544     $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate
545     ellipsoid about its long axis approaches $0$, as no solvent is
546     displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for
547     rotation of the prolate ellipsoid about its short axis is $35.9\%$ of
548     the analytical $\Xi^{rr}_{\mathit{stick}}$ result, accounting for the
549     reduced interfacial friction under ``slip'' boundary conditions.
550 kstocke1 3991
551 kstocke1 4058
552 gezelter 4064 An
553     $\eta$ value for TraPPE-UA hexane under these particular temperature
554     and pressure conditions was determined by applying a traditional
555     VSS-RNEMD linear momentum flux to a periodic box of solvent.
556    
557 kstocke1 4058 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 ($\omega_{Au}$)
558    
559 kstocke1 3947 \begin{equation}
560 kstocke1 3991 \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
561 kstocke1 4003 \label{eq:Xieff}
562 kstocke1 3947 \end{equation}
563    
564 kstocke1 3991 The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
565    
566     \begin{equation}
567     \tau = \frac{L}{2 t}
568 kstocke1 4003 \label{eq:tau}
569 kstocke1 3991 \end{equation}
570    
571     where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
572    
573 kstocke1 3947 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
574 kstocke1 3927 % **TESTS AND APPLICATIONS**
575     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
576     \section{Tests and Applications}
577    
578     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
579     % THERMAL CONDUCTIVITIES
580     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
581     \subsection{Thermal conductivities}
582    
583 kstocke1 4009 Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at
584     different kinetic energy flux values are shown in Table \ref{table:goldTC}. For these calculations, the hot and
585     cold slabs were excluded from the linear regression of the thermal gradient.
586 kstocke1 3927
587 kstocke1 3934 \begin{longtable}{ccc}
588     \caption{Calculated thermal conductivity of a crystalline gold nanoparticle of radius 40 \AA. Calculations were performed at 300 K and ambient density. Gold-gold interactions are described by the Quantum Sutton-Chen potential.}
589 kstocke1 3927 \\ \hline \hline
590 kstocke1 4003 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
591     {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
592 kstocke1 3991 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
593     6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
594     1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
595     3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
596     6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
597     \hline
598     This work & & 1.0040
599 kstocke1 3934 \\ \hline \hline
600 kstocke1 3991 \label{table:goldTC}
601 kstocke1 3927 \end{longtable}
602    
603 kstocke1 4009 The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly dependent on the applied kinetic energy
604 gezelter 4063 flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 -- 1.26 W / m $\cdot$ K\cite{Kuang:2010if}, though still significantly lower than the experimental value
605 kstocke1 4058 of 320 W / m $\cdot$ K, as the QSC force field neglects significant electronic contributions to
606 kstocke1 4009 heat conduction.
607 kstocke1 3947
608 kstocke1 4009 Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table
609     \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were
610     excluded from the $\langle \frac{dT}{dr} \rangle$ fit.
611    
612 kstocke1 3934 \begin{longtable}{ccc}
613 kstocke1 3991 \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
614 kstocke1 3927 \\ \hline \hline
615 kstocke1 4003 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
616     {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
617 kstocke1 3991 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
618     3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
619     6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
620     \hline
621     This work & & 0.8841 \\
622     Zhang, et al\cite{Zhang2005} & & 0.81 \\
623     R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
624     Experiment\cite{WagnerKruse} & & 0.61
625     \\ \hline \hline
626     \label{table:waterTC}
627 kstocke1 3927 \end{longtable}
628    
629 kstocke1 4009 Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average
630 kstocke1 4058 calculated thermal conductivity from this work, $0.8841$ W / m $\cdot$ K, compares very well to
631 kstocke1 4009 previous non-equilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental
632     values.\cite{WagnerKruse}
633    
634 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
635     % INTERFACIAL THERMAL CONDUCTANCE
636     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
637     \subsection{Interfacial thermal conductance}
638    
639 kstocke1 4009 Calculated interfacial thermal conductance ($G$) values for three sizes of gold nanoparticles and Au(111)
640     surface solvated in TraPPE-UA hexane are shown in Table \ref{table:G}.
641 kstocke1 4003
642 kstocke1 3973 \begin{longtable}{ccc}
643 kstocke1 4058 \caption{Calculated interfacial thermal conductance ($G$) values for gold nanoparticles of varying radii solvated in TraPPE-UA hexane. The nanoparticle $G$ values are compared to previous simulation results for a Au(111) interface in TraPPE-UA hexane.}
644 kstocke1 3973 \\ \hline \hline
645 kstocke1 4003 {Nanoparticle Radius} & {$G$}\\
646     {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
647     20 & {47.1} \\
648     30 & {45.4} \\
649     40 & {46.5} \\
650 kstocke1 4004 \hline
651     Au(111) & {30.2}
652 kstocke1 4003 \\ \hline \hline
653 kstocke1 4009 \label{table:G}
654 kstocke1 3973 \end{longtable}
655 kstocke1 3962
656 kstocke1 4009 The introduction of surface curvature increases the interfacial thermal conductance by a factor of
657     approximately $1.5$ relative to the flat interface. There are no significant differences in the $G$ values for
658     the varying nanoparticle sizes. It seems likely that for the range of nanoparticle sizes represented here, any
659 kstocke1 4058 particle size effects are not evident. The simulation of larger nanoparticles may demonstrate an approach to the $G$ value of a flat Au(111) slab but would require prohibitively costly numbers of atoms.
660 kstocke1 4009
661 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
662     % INTERFACIAL FRICTION
663     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
664     \subsection{Interfacial friction}
665    
666 kstocke1 4009 Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold
667     nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied
668     between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius,
669     respectively.
670 kstocke1 4004
671 kstocke1 4003 \begin{longtable}{lccccc}
672 kstocke1 4009 \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.}
673 kstocke1 3927 \\ \hline \hline
674 kstocke1 4003 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
675     {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
676     Sphere (r = 20 \AA) & {$x = y = z$} & 0 & {2386} & {3314} & {0.720}\\
677     Sphere (r = 30 \AA) & {$x = y = z$} & 0 & {8415} & {11749} & {0.716}\\
678     Sphere (r = 40 \AA) & {$x = y = z$} & 0 & {47544} & {34464} & {1.380}\\
679     Prolate Ellipsoid & {$x = y$} & {1792} & {3128} & {4991} & {0.627}\\
680 kstocke1 4004 Prolate Ellipsoid & {$z$} & 0 & {1590} & {1993} & {0.798}
681 kstocke1 4003 \\ \hline \hline
682 kstocke1 3991 \label{table:couple}
683 kstocke1 3927 \end{longtable}
684    
685 kstocke1 4009 The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold
686 kstocke1 4058 structures solvated by hexane do not exist in the ``slip'' boundary condition. At this length scale, the
687 kstocke1 4009 nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience
688     increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius
689     nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of
690     the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short
691 kstocke1 4058 axis, which at first seems counterintuitive. However, the `propellor' motion caused by rotation about the
692 kstocke1 4009 short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the
693     gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius)
694     appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of
695     surface features or anomalous solvent behaviors that are not fully understood at this time.
696 kstocke1 3994
697 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
698     % **DISCUSSION**
699     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
700     \section{Discussion}
701    
702 kstocke1 4009 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.
703 kstocke1 3927
704 kstocke1 4010 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.
705 kstocke1 4009
706 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
707     % **ACKNOWLEDGMENTS**
708     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
709 gezelter 4063 \begin{acknowledgement}
710     The authors thank Dr. Shenyu Kuang for helpful discussions. Support
711     for this project was provided by the National Science Foundation
712     under grant CHE-0848243. Computational time was provided by the
713     Center for Research Computing (CRC) at the University of Notre Dame.
714     \end{acknowledgement}
715 kstocke1 3927
716    
717     \newpage
718    
719     \bibliography{nonperiodicVSS}
720    
721 gezelter 4063 %\end{doublespace}
722 kstocke1 3934 \end{document}

Properties

Name Value
svn:executable