ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3779
Committed: Sat Dec 10 23:46:08 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 36080 byte(s)
Log Message:
a rough version of first draft!

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