ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3784
Committed: Mon Dec 12 23:39:21 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 37076 byte(s)
Log Message:
add new reference and some edits in simulation details.

File Contents

# User Rev Content
1 gezelter 3769 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5     \usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{tabularx}
8     \usepackage{graphicx}
9     \usepackage{multirow}
10     %\usepackage{booktabs}
11     %\usepackage{bibentry}
12     %\usepackage{mathrsfs}
13     %\usepackage[ref]{overcite}
14     \usepackage[square, comma, sort&compress]{natbib}
15     \usepackage{url}
16     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19    
20     % double space list of tables and figures
21     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22     \setlength{\abovecaptionskip}{20 pt}
23     \setlength{\belowcaptionskip}{30 pt}
24    
25     %\renewcommand\citemid{\ } % no comma in optional reference note
26     \bibpunct{[}{]}{,}{n}{}{;}
27     \bibliographystyle{achemso}
28    
29     \begin{document}
30    
31 skuang 3770 \title{ENTER TITLE HERE}
32 gezelter 3769
33     \author{Shenyu Kuang and J. Daniel
34     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
35     Department of Chemistry and Biochemistry,\\
36     University of Notre Dame\\
37     Notre Dame, Indiana 46556}
38    
39     \date{\today}
40    
41     \maketitle
42    
43     \begin{doublespace}
44    
45     \begin{abstract}
46 skuang 3779 We present a new method for introducing stable nonequilibrium
47     velocity and temperature gradients in molecular dynamics simulations
48     of heterogeneous systems. This method conserves the linear momentum
49     and total energy of the system and improves previous Reverse
50     Non-Equilibrium Molecular Dynamics (RNEMD) methods and maintains
51     thermal velocity distributions. It also avoid thermal anisotropy
52     occured in NIVS simulations by using isotropic velocity scaling on
53     the molecules in specific regions of a system. To test the method,
54     we have computed the thermal conductivity and shear viscosity of
55     model liquid systems as well as the interfacial frictions of a
56     series of metal/liquid interfaces.
57 gezelter 3769
58     \end{abstract}
59    
60     \newpage
61    
62     %\narrowtext
63    
64     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65     % BODY OF TEXT
66     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67    
68     \section{Introduction}
69 skuang 3771 [REFINE LATER, ADD MORE REF.S]
70     Imposed-flux methods in Molecular Dynamics (MD)
71     simulations\cite{MullerPlathe:1997xw} can establish steady state
72     systems with a set applied flux vs a corresponding gradient that can
73     be measured. These methods does not need many trajectories to provide
74     information of transport properties of a given system. Thus, they are
75     utilized in computing thermal and mechanical transfer of homogeneous
76     or bulk systems as well as heterogeneous systems such as liquid-solid
77     interfaces.\cite{kuang:AuThl}
78 skuang 3770
79 skuang 3771 The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that
80     satisfy linear momentum and total energy conservation of a system when
81     imposing fluxes in a simulation. Thus they are compatible with various
82     ensembles, including the micro-canonical (NVE) ensemble, without the
83     need of an external thermostat. The original approaches by
84     M\"{u}ller-Plathe {\it et
85     al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
86     momentum swapping for generating energy/momentum fluxes, which is also
87     compatible with particles of different identities. Although simple to
88     implement in a simulation, this approach can create nonthermal
89     velocity distributions, as discovered by Tenney and
90     Maginn\cite{Maginn:2010}. Furthermore, this approach to kinetic energy
91     transfer between particles of different identities is less efficient
92     when the mass difference between the particles becomes significant,
93     which also limits its application on heterogeneous interfacial
94     systems.
95 skuang 3770
96 skuang 3771 Recently, we developed a different approach, using Non-Isotropic
97     Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose
98     fluxes. Compared to the momentum swapping move, it scales the velocity
99     vectors in two separate regions of a simulated system with respective
100     diagonal scaling matrices. These matrices are determined by solving a
101     set of equations including linear momentum and kinetic energy
102     conservation constraints and target flux satisfaction. This method is
103     able to effectively impose a wide range of kinetic energy fluxes
104     without obvious perturbation to the velocity distributions of the
105     simulated systems, regardless of the presence of heterogeneous
106     interfaces. We have successfully applied this approach in studying the
107     interfacial thermal conductance at metal-solvent
108     interfaces.\cite{kuang:AuThl}
109 gezelter 3769
110 skuang 3771 However, the NIVS approach limits its application in imposing momentum
111     fluxes. Temperature anisotropy can happen under high momentum fluxes,
112     due to the nature of the algorithm. Thus, combining thermal and
113     momentum flux is also difficult to implement with this
114     approach. However, such combination may provide a means to simulate
115     thermal/momentum gradient coupled processes such as freeze
116     desalination. Therefore, developing novel approaches to extend the
117     application of imposed-flux method is desired.
118 gezelter 3769
119 skuang 3771 In this paper, we improve the NIVS method and propose a novel approach
120     to impose fluxes. This approach separate the means of applying
121     momentum and thermal flux with operations in one time step and thus is
122     able to simutaneously impose thermal and momentum flux. Furthermore,
123     the approach retains desirable features of previous RNEMD approaches
124     and is simpler to implement compared to the NIVS method. In what
125     follows, we first present the method to implement the method in a
126     simulation. Then we compare the method on bulk fluids to previous
127     methods. Also, interfacial frictions are computed for a series of
128     interfaces.
129 gezelter 3769
130     \section{Methodology}
131 skuang 3781 Similar to the NIVS method,\cite{kuang:164101} we consider a
132 skuang 3770 periodic system divided into a series of slabs along a certain axis
133     (e.g. $z$). The unphysical thermal and/or momentum flux is designated
134 skuang 3781 from the center slab to one of the end slabs, and thus the thermal
135     flux results in a lower temperature of the center slab than the end
136     slab, and the momentum flux results in negative center slab momentum
137     with positive end slab momentum (unless these fluxes are set
138     negative). Therefore, the center slab is denoted as ``$c$'', while the
139     end slab as ``$h$''.
140 skuang 3770
141 skuang 3781 To impose these fluxes, we periodically apply different set of
142     operations on velocities of particles {$i$} within the center slab and
143     those of particles {$j$} within the end slab:
144 skuang 3770 \begin{eqnarray}
145     \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
146     \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
147     \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
148     \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
149     \end{eqnarray}
150     where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes
151     the instantaneous bulk velocity of slabs $c$ and $h$ respectively
152 skuang 3781 before an operation is applied. When a momentum flux $\vec{j}_z(\vec{p})$
153 skuang 3770 presents, these bulk velocities would have a corresponding change
154     ($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's
155     second law:
156     \begin{eqnarray}
157     M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\
158     M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
159     \end{eqnarray}
160 skuang 3781 where $M$ denotes total mass of particles within a slab:
161 skuang 3770 \begin{eqnarray}
162     M_c & = & \sum_{i = 1}^{N_c} m_i \\
163     M_h & = & \sum_{j = 1}^{N_h} m_j
164     \end{eqnarray}
165 skuang 3781 and $\Delta t$ is the interval between two separate operations.
166 skuang 3770
167 skuang 3781 The above operations already conserve the linear momentum of a
168     periodic system. To further satisfy total energy conservation as well
169     as to impose the thermal flux $J_z$, the following equations are
170     included as well:
171 skuang 3784 [MAY PUT EXTRA MATH IN SUPPORT INFO OR APPENDIX]
172 skuang 3770 \begin{eqnarray}
173     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
174     \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
175     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
176     \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2
177     \end{eqnarray}
178     where $K_c$ and $K_h$ denotes translational kinetic energy of slabs
179 skuang 3781 $c$ and $h$ respectively before an operation is applied. These
180 skuang 3770 translational kinetic energy conservation equations are sufficient to
181 skuang 3781 ensure total energy conservation, as the operations applied in our
182     method do not change the kinetic energy related to other degrees of
183     freedom or the potential energy of a system, given that its potential
184 skuang 3770 energy does not depend on particle velocity.
185    
186     The above sets of equations are sufficient to determine the velocity
187     scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
188 skuang 3781 $\vec{a}_h$. Note that there are two roots respectively for $c$ and
189     $h$. However, the positive roots (which are closer to 1) are chosen so
190     that the perturbations to a system can be reduced to a minimum. Figure
191     \ref{method} illustrates the implementation sketch of this algorithm
192     in an individual step.
193 skuang 3770
194 skuang 3772 \begin{figure}
195     \includegraphics[width=\linewidth]{method}
196     \caption{Illustration of the implementation of the algorithm in a
197     single step. Starting from an ideal velocity distribution, the
198 skuang 3781 transformation is used to apply the effect of both a thermal and a
199     momentum flux from the ``c'' slab to the ``h'' slab. As the figure
200     shows, thermal distributions can preserve after this operation.}
201 skuang 3772 \label{method}
202     \end{figure}
203    
204 skuang 3770 By implementing these operations at a certain frequency, a steady
205     thermal and/or momentum flux can be applied and the corresponding
206     temperature and/or momentum gradients can be established.
207    
208 skuang 3781 Compared to the previous NIVS method, this approach is computationally
209     more efficient in that only quadratic equations are involved to
210     determine a set of scaling coefficients, while the NIVS method needs
211     to solve quartic equations. Furthermore, this method implements
212     isotropic scaling of velocities in respective slabs, unlike the NIVS,
213     where an extra criteria function is necessary to choose a set of
214     coefficients that performs a scaling as isotropic as possible. More
215     importantly, separating the means of momentum flux imposing from
216     velocity scaling avoids the underlying cause to thermal anisotropy in
217     NIVS when applying a momentum flux. And later sections will
218     demonstrate that this can improve the performance in shear viscosity
219     simulations.
220 gezelter 3769
221 skuang 3781 The advantages of this approach over the original momentum swapping
222     one lies in its nature of preserve a Maxwell-Boltzmann distribution
223     (mathimatically a Gaussian function). However, because the momentum
224     swapping steps tend to result in a nonthermal distribution, when an
225     imposed flux is relatively large, diffusions from the neighboring
226     slabs could no longer remedy this effect, problematic distributions
227     would be observed. Results in later section will illustrate this
228     effect in more detail.
229 gezelter 3769
230 skuang 3772 \section{Computational Details}
231 skuang 3773 The algorithm has been implemented in our MD simulation code,
232     OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with
233 skuang 3781 previous RNEMD methods or equilibrium MD (EMD) methods in homogeneous fluids
234 skuang 3773 (Lennard-Jones and SPC/E water). And taking advantage of the method,
235     we simulate the interfacial friction of different heterogeneous
236     interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid
237     water).
238 gezelter 3769
239 skuang 3773 \subsection{Simulation Protocols}
240 skuang 3784 The systems to be investigated are set up in orthorhombic simulation
241     cells with periodic boundary conditions in all three dimensions. The
242     $z$ axis of these cells were longer and set as the temperature and/or
243     momentum gradient axis. And the cells were evenly divided into $N$
244 skuang 3773 slabs along this axis, with various $N$ depending on individual
245 skuang 3784 system. The $x$ and $y$ axis were of the same length in homogeneous
246     systems or had length scale close to each other where heterogeneous
247     interfaces presents. In all cases, before introducing a nonequilibrium
248     method to establish steady thermal and/or momentum gradients for
249     further measurements and calculations, canonical ensemble with a
250     Nos\'e-Hoover thermostat\cite{hoover85} and microcanonical ensemble
251     equilibrations were used before data collections. For SPC/E water
252     simulations, isobaric-isothermal equilibrations\cite{melchionna93} are
253     performed before the above to reach normal pressure (1 bar); for
254     interfacial systems, similar equilibrations are used to relax the
255     surface tensions of the $xy$ plane.
256 skuang 3772
257 skuang 3784 While homogeneous fluid systems can be set up with rather random
258     configurations, our interfacial systems needs a series of steps to
259     ensure the interfaces be established properly for computations. The
260 skuang 3774 preparation and equilibration of butanethiol covered gold (111)
261     surface and further solvation and equilibration process is described
262 skuang 3784 in details as in reference \cite{kuang:AuThl}.
263 skuang 3773
264 skuang 3774 As for the ice/liquid water interfaces, the basal surface of ice
265 skuang 3775 lattice was first constructed. Hirsch {\it et
266     al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice
267 skuang 3784 lattices with all possible proton order configurations. We refer to
268     their results and choose the configuration of the lowest energy after
269     geometry optimization as the unit cell for our ice lattices. Although
270     experimental solid/liquid coexistant temperature under normal pressure
271     should be close to 273K, Bryk and Haymet's simulations of ice/liquid
272     water interfaces with different models suggest that for SPC/E, the
273     most stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
274     Therefore, our ice/liquid water simulations were carried out at
275     225K. To have extra protection of the ice lattice during initial
276     equilibration (when the randomly generated liquid phase configuration
277     could release large amount of energy in relaxation), restraints were
278     applied to the ice lattice to avoid inadvertent melting by the heat
279     dissipated from the high enery configurations.
280     [MAY ADD A SNAPSHOT FOR BASAL PLANE]
281 gezelter 3769
282     \subsection{Force Field Parameters}
283 skuang 3774 For comparison of our new method with previous work, we retain our
284 skuang 3784 force field parameters consistent with previous simulations. Argon is
285     the Lennard-Jones fluid used here, and its results are reported in
286     reduced unit for direct comparison purpose.
287 gezelter 3769
288 skuang 3774 As for our water simulations, SPC/E model is used throughout this work
289     for consistency. Previous work for transport properties of SPC/E water
290 skuang 3775 model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
291     that unnecessary repetition of previous methods can be avoided.
292 gezelter 3769
293 skuang 3774 The Au-Au interaction parameters in all simulations are described by
294     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
295     QSC potentials include zero-point quantum corrections and are
296 gezelter 3769 reparametrized for accurate surface energies compared to the
297 skuang 3775 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
298     Spohr potential was adopted\cite{ISI:000167766600035} to depict
299     Au-H$_2$O interactions.
300 gezelter 3769
301 skuang 3784 For our gold/organic liquid interfaces, the small organic molecules
302     included in our simulations are the Au surface capping agent
303     butanethiol and liquid hexane and toluene. The United-Atom
304 skuang 3774 models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
305     for these components were used in this work for better computational
306     efficiency, while maintaining good accuracy. We refer readers to our
307     previous work\cite{kuang:AuThl} for further details of these models,
308     as well as the interactions between Au and the above organic molecule
309     components.
310 gezelter 3769
311 skuang 3774 \subsection{Thermal conductivities}
312 skuang 3775 When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
313     impose kinetic energy transfer, the method can be used for thermal
314     conductivity computations. Similar to previous RNEMD methods, we
315     assume linear response of the temperature gradient with respect to the
316     thermal flux in general case. And the thermal conductivity ($\lambda$)
317     can be obtained with the imposed kinetic energy flux and the measured
318     thermal gradient:
319     \begin{equation}
320     J_z = -\lambda \frac{\partial T}{\partial z}
321     \end{equation}
322     Like other imposed-flux methods, the energy flux was calculated using
323     the total non-physical energy transferred (${E_{total}}$) from slab
324     ``c'' to slab ``h'', which is recorded throughout a simulation, and
325     the time for data collection $t$:
326     \begin{equation}
327     J_z = \frac{E_{total}}{2 t L_x L_y}
328     \end{equation}
329     where $L_x$ and $L_y$ denotes the dimensions of the plane in a
330     simulation cell perpendicular to the thermal gradient, and a factor of
331 skuang 3784 two in the denominator is necessary for the heat transport occurs in
332     both $+z$ and $-z$ directions. The average temperature gradient
333 skuang 3775 ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
334     regression of the temperature profile, which is recorded during a
335     simulation for each slab in a cell. For Lennard-Jones simulations,
336     thermal conductivities are reported in reduced units
337     (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
338    
339 skuang 3774 \subsection{Shear viscosities}
340 skuang 3775 Alternatively, the method can carry out shear viscosity calculations
341 skuang 3784 by specify a momentum flux. In our algorithm, one can specify the
342     three components of the flux vector $\vec{j}_z(\vec{p})$
343 skuang 3775 respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
344 skuang 3784 set to zero. For isotropic systems, the direction of
345     $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
346     ability of arbitarily specifying the vector direction in our method
347     could provide convenience in anisotropic simulations.
348 skuang 3775
349 skuang 3784 Similar to thermal conductivity computations, for a homogeneous
350     system, linear response of the momentum gradient with respect to the
351     shear stress is assumed, and the shear viscosity ($\eta$) can be
352     obtained with the imposed momentum flux (e.g. in $x$ direction) and
353     the measured gradient:
354 skuang 3775 \begin{equation}
355     j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
356     \end{equation}
357     where the flux is similarly defined:
358     \begin{equation}
359     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
360     \end{equation}
361     with $P_x$ being the total non-physical momentum transferred within
362 skuang 3784 the data collection time. Also, the averaged velocity gradient
363 skuang 3775 ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
364 skuang 3784 regression of the $x$ component of the mean velocity ($\langle
365     v_x\rangle$) in each of the bins. For Lennard-Jones simulations, shear
366     viscosities are also reported in reduced units
367 skuang 3775 (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
368    
369 skuang 3784 [COMBINE JZKE W JZPX]
370    
371 skuang 3774 \subsection{Interfacial friction and Slip length}
372 skuang 3775 While the shear stress results in a velocity gradient within bulk
373     fluid phase, its effect at a solid-liquid interface could vary due to
374     the interaction strength between the two phases. The interfacial
375     friction coefficient $\kappa$ is defined to relate the shear stress
376     (e.g. along $x$-axis) and the relative fluid velocity tangent to the
377     interface:
378     \begin{equation}
379     j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
380     \end{equation}
381     Under ``stick'' boundary condition, $\Delta v_x|_{interface}
382     \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
383     ``slip'' boundary condition at the solid-liquid interface, $\kappa$
384     becomes finite. To characterize the interfacial boundary conditions,
385     slip length ($\delta$) is defined using $\kappa$ and the shear
386     viscocity of liquid phase ($\eta$):
387     \begin{equation}
388     \delta = \frac{\eta}{\kappa}
389     \end{equation}
390     so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
391     and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
392     illustrates how this quantity is defined and computed for a
393 skuang 3776 solid-liquid interface. [MAY INCLUDE A SNAPSHOT IN FIGURE]
394 gezelter 3769
395 skuang 3775 \begin{figure}
396     \includegraphics[width=\linewidth]{defDelta}
397     \caption{The slip length $\delta$ can be obtained from a velocity
398 skuang 3780 profile of a solid-liquid interface simulation. An example of
399     Au/hexane interfaces is shown. Calculation for the left side is
400     illustrated. The right side is similar to the left side.}
401 skuang 3775 \label{slipLength}
402     \end{figure}
403 gezelter 3769
404 skuang 3775 In our method, a shear stress can be applied similar to shear
405     viscosity computations by applying an unphysical momentum flux
406     (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
407     shown in Figure \ref{slipLength}, in which the velocity gradients
408     within liquid phase and velocity difference at the liquid-solid
409     interface can be measured respectively. Further calculations and
410     characterizations of the interface can be carried out using these
411     data.
412    
413 skuang 3776 \section{Results and Discussions}
414     \subsection{Lennard-Jones fluid}
415     Our orthorhombic simulation cell of Lennard-Jones fluid has identical
416     parameters to our previous work\cite{kuang:164101} to facilitate
417     comparison. Thermal conductivitis and shear viscosities were computed
418     with the algorithm applied to the simulations. The results of thermal
419     conductivity are compared with our previous NIVS algorithm. However,
420     since the NIVS algorithm could produce temperature anisotropy for
421     shear viscocity computations, these results are instead compared to
422     the momentum swapping approaches. Table \ref{LJ} lists these
423     calculations with various fluxes in reduced units.
424 skuang 3770
425 skuang 3776 \begin{table*}
426     \begin{minipage}{\linewidth}
427     \begin{center}
428 gezelter 3769
429 skuang 3776 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
430     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
431     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
432     at various momentum fluxes. The new method yields similar
433     results to previous RNEMD methods. All results are reported in
434     reduced unit. Uncertainties are indicated in parentheses.}
435    
436     \begin{tabular}{cccccc}
437     \hline\hline
438     \multicolumn{2}{c}{Momentum Exchange} &
439     \multicolumn{2}{c}{$\lambda^*$} &
440     \multicolumn{2}{c}{$\eta^*$} \\
441     \hline
442 skuang 3779 Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
443 skuang 3784 NIVS\footnote{Reference \cite{kuang:164101}.} & This work &
444     Swapping & This work \\
445 skuang 3776 \hline
446     0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
447     0.232 & 0.09 & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
448     0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
449     0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
450     1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
451     \hline\hline
452     \end{tabular}
453     \label{LJ}
454     \end{center}
455     \end{minipage}
456     \end{table*}
457 gezelter 3769
458 skuang 3776 \subsubsection{Thermal conductivity}
459     Our thermal conductivity calculations with this method yields
460     comparable results to the previous NIVS algorithm. This indicates that
461     the thermal gradients rendered using this method are also close to
462     previous RNEMD methods. Simulations with moderately higher thermal
463     fluxes tend to yield more reliable thermal gradients and thus avoid
464     large errors, while overly high thermal fluxes could introduce side
465     effects such as non-linear temperature gradient response or
466     inadvertent phase transitions.
467 gezelter 3769
468 skuang 3776 Since the scaling operation is isotropic in this method, one does not
469     need extra care to ensure temperature isotropy between the $x$, $y$
470     and $z$ axes, while thermal anisotropy might happen if the criteria
471     function for choosing scaling coefficients does not perform as
472     expected. Furthermore, this method avoids inadvertent concomitant
473     momentum flux when only thermal flux is imposed, which could not be
474     achieved with swapping or NIVS approaches. The thermal energy exchange
475 skuang 3778 in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
476 skuang 3776 or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
477     P^\alpha$) would not obtain this result unless thermal flux vanishes
478     (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which are trivial to apply a
479     thermal flux). In this sense, this method contributes to having
480     minimal perturbation to a simulation while imposing thermal flux.
481    
482     \subsubsection{Shear viscosity}
483     Table \ref{LJ} also compares our shear viscosity results with momentum
484     swapping approach. Our calculations show that our method predicted
485     similar values for shear viscosities to the momentum swapping
486     approach, as well as the velocity gradient profiles. Moderately larger
487     momentum fluxes are helpful to reduce the errors of measured velocity
488     gradients and thus the final result. However, it is pointed out that
489     the momentum swapping approach tends to produce nonthermal velocity
490     distributions.\cite{Maginn:2010}
491    
492     To examine that temperature isotropy holds in simulations using our
493     method, we measured the three one-dimensional temperatures in each of
494     the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
495     temperatures were calculated after subtracting the effects from bulk
496     velocities of the slabs. The one-dimensional temperature profiles
497     showed no observable difference between the three dimensions. This
498     ensures that isotropic scaling automatically preserves temperature
499     isotropy and that our method is useful in shear viscosity
500     computations.
501    
502 gezelter 3769 \begin{figure}
503 skuang 3776 \includegraphics[width=\linewidth]{tempXyz}
504 skuang 3777 \caption{Unlike the previous NIVS algorithm, the new method does not
505     produce a thermal anisotropy. No temperature difference between
506     different dimensions were observed beyond the magnitude of the error
507     bars. Note that the two ``hotter'' regions are caused by the shear
508     stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
509     an effect that only observed in our methods.}
510 skuang 3776 \label{tempXyz}
511 gezelter 3769 \end{figure}
512    
513 skuang 3776 Furthermore, the velocity distribution profiles are tested by imposing
514     a large shear stress into the simulations. Figure \ref{vDist}
515     demonstrates how our method is able to maintain thermal velocity
516     distributions against the momentum swapping approach even under large
517     imposed fluxes. Previous swapping methods tend to deplete particles of
518     positive velocities in the negative velocity slab (``c'') and vice
519     versa in slab ``h'', where the distributions leave a notch. This
520     problematic profiles become significant when the imposed-flux becomes
521     larger and diffusions from neighboring slabs could not offset the
522     depletion. Simutaneously, abnormal peaks appear corresponding to
523     excessive velocity swapped from the other slab. This nonthermal
524     distributions limit applications of the swapping approach in shear
525     stress simulations. Our method avoids the above problematic
526     distributions by altering the means of applying momentum
527     fluxes. Comparatively, velocity distributions recorded from
528     simulations with our method is so close to the ideal thermal
529     prediction that no observable difference is shown in Figure
530     \ref{vDist}. Conclusively, our method avoids problems happened in
531     previous RNEMD methods and provides a useful means for shear viscosity
532     computations.
533 gezelter 3769
534 skuang 3776 \begin{figure}
535     \includegraphics[width=\linewidth]{velDist}
536 skuang 3777 \caption{Velocity distributions that develop under the swapping and
537     our methods at high flux. These distributions were obtained from
538 skuang 3779 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
539 skuang 3777 swapping interval of 20 time steps). This is a relatively large flux
540     to demonstrate the nonthermal distributions that develop under the
541     swapping method. Distributions produced by our method are very close
542     to the ideal thermal situations.}
543 skuang 3776 \label{vDist}
544     \end{figure}
545 gezelter 3769
546 skuang 3776 \subsection{Bulk SPC/E water}
547 skuang 3778 Since our method was in good performance of thermal conductivity and
548     shear viscosity computations for simple Lennard-Jones fluid, we extend
549     our applications of these simulations to complex fluid like SPC/E
550     water model. A simulation cell with 1000 molecules was set up in the
551     same manner as in \cite{kuang:164101}. For thermal conductivity
552     simulations, measurements were taken to compare with previous RNEMD
553     methods; for shear viscosity computations, simulations were run under
554     a series of temperatures (with corresponding pressure relaxation using
555     the isobaric-isothermal ensemble[CITE NIVS REF 32]), and results were
556     compared to available data from Equilibrium MD methods[CITATIONS].
557 skuang 3777
558 skuang 3776 \subsubsection{Thermal conductivity}
559 skuang 3778 Table \ref{spceThermal} summarizes our thermal conductivity
560     computations under different temperatures and thermal gradients, in
561     comparison to the previous NIVS results\cite{kuang:164101} and
562     experimental measurements\cite{WagnerKruse}. Note that no appreciable
563     drift of total system energy or temperature was observed when our
564     method is applied, which indicates that our algorithm conserves total
565     energy even for systems involving electrostatic interactions.
566 gezelter 3769
567 skuang 3778 Measurements using our method established similar temperature
568     gradients to the previous NIVS method. Our simulation results are in
569     good agreement with those from previous simulations. And both methods
570     yield values in reasonable agreement with experimental
571     values. Simulations using moderately higher thermal gradient or those
572     with longer gradient axis ($z$) for measurement seem to have better
573     accuracy, from our results.
574    
575     \begin{table*}
576     \begin{minipage}{\linewidth}
577     \begin{center}
578    
579     \caption{Thermal conductivity of SPC/E water under various
580     imposed thermal gradients. Uncertainties are indicated in
581     parentheses.}
582    
583     \begin{tabular}{ccccc}
584     \hline\hline
585     $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
586     {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
587     (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
588     Experiment\cite{WagnerKruse} \\
589     \hline
590     300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\
591     318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
592     & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
593 skuang 3779 & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
594     twice as long.} & & \\
595 skuang 3778 \hline\hline
596     \end{tabular}
597     \label{spceThermal}
598     \end{center}
599     \end{minipage}
600     \end{table*}
601    
602 skuang 3776 \subsubsection{Shear viscosity}
603 skuang 3778 The improvement our method achieves for shear viscosity computations
604     enables us to apply it on SPC/E water models. The series of
605     temperatures under which our shear viscosity calculations were carried
606     out covers the liquid range under normal pressure. Our simulations
607     predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
608 skuang 3779 (Table \ref{spceShear}). Considering subtlties such as temperature or
609     pressure/density errors in these two series of measurements, our
610     results show no significant difference from those with EMD
611     methods. Since each value reported using our method takes only one
612     single trajectory of simulation, instead of average from many
613     trajectories when using EMD, our method provides an effective means
614     for shear viscosity computations.
615 gezelter 3769
616 skuang 3778 \begin{table*}
617     \begin{minipage}{\linewidth}
618     \begin{center}
619    
620     \caption{Computed shear viscosity of SPC/E water under different
621     temperatures. Results are compared to those obtained with EMD
622     method[CITATION]. Uncertainties are indicated in parentheses.}
623    
624     \begin{tabular}{cccc}
625     \hline\hline
626     $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
627     {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
628     (K) & (10$^{10}$s$^{-1}$) & This work & Previous simulations[CITATION]\\
629     \hline
630     273 & & 1.218(0.004) & \\
631     & & 1.140(0.012) & \\
632     303 & & 0.646(0.008) & \\
633     318 & & 0.536(0.007) & \\
634     & & 0.510(0.007) & \\
635     & & & \\
636     333 & & 0.428(0.002) & \\
637     363 & & 0.279(0.014) & \\
638     & & 0.306(0.001) & \\
639     \hline\hline
640     \end{tabular}
641     \label{spceShear}
642     \end{center}
643     \end{minipage}
644     \end{table*}
645 gezelter 3769
646 skuang 3776 [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
647 skuang 3777 [PUT RESULTS AND FIGURE HERE IF IT WORKS]
648 skuang 3779 \subsection{Interfacial frictions and slip lengths}
649     An attractive aspect of our method is the ability to apply momentum
650     and/or thermal flux in nonhomogeneous systems, where molecules of
651     different identities (or phases) are segregated in different
652     regions. We have previously studied the interfacial thermal transport
653     of a series of metal gold-liquid
654     surfaces\cite{kuang:164101,kuang:AuThl}, and attemptions have been
655     made to investigate the relationship between this phenomenon and the
656     interfacial frictions.
657 gezelter 3769
658 skuang 3779 Table \ref{etaKappaDelta} includes these computations and previous
659     calculations of corresponding interfacial thermal conductance. For
660     bare Au(111) surfaces, slip boundary conditions were observed for both
661     organic and aqueous liquid phases, corresponding to previously
662     computed low interfacial thermal conductance. Instead, the butanethiol
663     covered Au(111) surface appeared to be sticky to the organic liquid
664     molecules in our simulations. We have reported conductance enhancement
665     effect for this surface capping agent,\cite{kuang:AuThl} and these
666     observations have a qualitative agreement with the thermal conductance
667     results. This agreement also supports discussions on the relationship
668     between surface wetting and slip effect and thermal conductance of the
669     interface.[CITE BARRAT, GARDE]
670 skuang 3776
671 gezelter 3769 \begin{table*}
672     \begin{minipage}{\linewidth}
673     \begin{center}
674    
675 skuang 3779 \caption{Computed interfacial friction coefficient values for
676     interfaces with various components for liquid and solid
677     phase. Error estimates are indicated in parentheses.}
678 gezelter 3769
679 skuang 3779 \begin{tabular}{llcccccc}
680 gezelter 3769 \hline\hline
681 skuang 3779 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
682     & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
683     \cite{kuang:164101}.} \\
684 skuang 3780 surface & molecules & K & MPa & mPa$\cdot$s & Pa$\cdot$s/m & nm
685     & MW/m$^2$/K \\
686 gezelter 3769 \hline
687 skuang 3779 Au(111) & hexane & 200 & 1.08 & 0.20() & 5.3$\times$10$^4$() &
688     3.7 & 46.5 \\
689     & & & 2.15 & 0.14() & 5.3$\times$10$^4$() &
690     2.7 & \\
691     Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.29() & $\infty$ & 0 &
692     131 \\
693     & & & 5.39 & 0.32() & $\infty$ & 0 &
694     \\
695 gezelter 3769 \hline
696 skuang 3779 Au(111) & toluene & 200 & 1.08 & 0.72() & 1.?$\times$10$^5$() &
697     4.6 & 70.1 \\
698     & & & 2.16 & 0.54() & 1.?$\times$10$^5$() &
699     4.9 & \\
700     Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.98() & $\infty$ & 0
701     & 187 \\
702     & & & 10.8 & 0.99() & $\infty$ & 0
703     & \\
704     \hline
705     Au(111) & water & 300 & 1.08 & 0.40() & 1.9$\times$10$^4$() &
706     20.7 & 1.65 \\
707     & & & 2.16 & 0.79() & 1.9$\times$10$^4$() &
708     41.9 & \\
709     \hline
710     ice(basal) & water & 225 & 19.4 & 15.8() & $\infty$ & 0 & \\
711 gezelter 3769 \hline\hline
712     \end{tabular}
713 skuang 3779 \label{etaKappaDelta}
714 gezelter 3769 \end{center}
715     \end{minipage}
716     \end{table*}
717    
718 skuang 3779 An interesting effect alongside the surface friction change is
719     observed on the shear viscosity of liquids in the regions close to the
720     solid surface. Note that $\eta$ measured near a ``slip'' surface tends
721     to be smaller than that near a ``stick'' surface. This suggests that
722     an interface could affect the dynamic properties on its neighbor
723     regions. It is known that diffusions of solid particles in liquid
724     phase is affected by their surface conditions (stick or slip
725     boundary).[CITE SCHMIDT AND SKINNER] Our observations could provide
726     support to this phenomenon.
727 gezelter 3769
728 skuang 3779 In addition to these previously studied interfaces, we attempt to
729     construct ice-water interfaces and the basal plane of ice lattice was
730     first studied. In contrast to the Au(111)/water interface, where the
731     friction coefficient is relatively small and large slip effect
732     presents, the ice/liquid water interface demonstrates strong
733     interactions and appears to be sticky. The supercooled liquid phase is
734     an order of magnitude viscous than measurements in previous
735     section. It would be of interst to investigate the effect of different
736     ice lattice planes (such as prism surface) on interfacial friction and
737     corresponding liquid viscosity.
738 gezelter 3769
739 skuang 3776 \section{Conclusions}
740 skuang 3779 Our simulations demonstrate the validity of our method in RNEMD
741     computations of thermal conductivity and shear viscosity in atomic and
742     molecular liquids. Our method maintains thermal velocity distributions
743     and avoids thermal anisotropy in previous NIVS shear stress
744     simulations, as well as retains attractive features of previous RNEMD
745     methods. There is no {\it a priori} restrictions to the method to be
746     applied in various ensembles, so prospective applications to
747     extended-system methods are possible.
748 gezelter 3769
749 skuang 3779 Furthermore, using this method, investigations can be carried out to
750     characterize interfacial interactions. Our method is capable of
751     effectively imposing both thermal and momentum flux accross an
752     interface and thus facilitates studies that relates dynamic property
753     measurements to the chemical details of an interface.
754 gezelter 3769
755 skuang 3779 Another attractive feature of our method is the ability of
756     simultaneously imposing thermal and momentum flux in a
757     system. potential researches that might be benefit include complex
758     systems that involve thermal and momentum gradients. For example, the
759     Soret effects under a velocity gradient would be of interest to
760     purification and separation researches.
761 gezelter 3769
762     \section{Acknowledgments}
763     Support for this project was provided by the National Science
764     Foundation under grant CHE-0848243. Computational time was provided by
765     the Center for Research Computing (CRC) at the University of Notre
766     Dame.
767    
768     \newpage
769    
770 skuang 3770 \bibliography{stokes}
771 gezelter 3769
772     \end{doublespace}
773     \end{document}