ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3781
Committed: Mon Dec 12 03:30:09 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 36804 byte(s)
Log Message:
some edits in methodology

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     [SUPPORT INFO OR APPENDIX MIGHT BE NECESSARY TO PUT EXTRA MATH IN]
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     The systems to be investigated are set up in a orthorhombic simulation
241     cell with periodic boundary conditions in all three dimensions. The
242 skuang 3774 $z$ axis of these cells were longer and was set as the gradient axis
243 skuang 3773 of temperature and/or momentum. Thus the cells were divided into $N$
244     slabs along this axis, with various $N$ depending on individual
245     system. The $x$ and $y$ axis were usually of the same length in
246     homogeneous systems or close to each other where interfaces
247     presents. In all cases, before introducing a nonequilibrium method to
248     establish steady thermal and/or momentum gradients for further
249     measurements and calculations, canonical ensemble with a Nos\'e-Hoover
250     thermostat\cite{hoover85} and microcanonical ensemble equilibrations
251     were used to prepare systems ready for data
252     collections. Isobaric-isothermal equilibrations are performed before
253     this for SPC/E water systems to reach normal pressure (1 bar), while
254     similar equilibrations are used for interfacial systems to relax the
255     surface tensions.
256 skuang 3772
257 skuang 3773 While homogeneous fluid systems can be set up with random
258     configurations, our interfacial systems needs extra steps to ensure
259 skuang 3774 the interfaces be established properly for computations. The
260     preparation and equilibration of butanethiol covered gold (111)
261     surface and further solvation and equilibration process is described
262 skuang 3775 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     lattices with different proton orders. We refer to their results and
268     choose the configuration of the lowest energy after geometry
269     optimization as the unit cells of our ice lattices. Although
270 skuang 3774 experimental solid/liquid coexistant temperature near normal pressure
271 skuang 3775 is 273K, Bryk and Haymet's simulations of ice/liquid water interfaces
272     with different models suggest that for SPC/E, the most stable
273     interface is observed at 225$\pm$5K. Therefore, all our ice/liquid
274     water simulations were carried out under 225K. To have extra
275     protection of the ice lattice during initial equilibration (when the
276     randomly generated liquid phase configuration could release large
277     amount of energy in relaxation), a constraint method (REF?) was
278     adopted until the high energy configuration was relaxed.
279     [MAY ADD A FIGURE HERE FOR BASAL PLANE, MAY INCLUDE PRISM IF POSSIBLE]
280 gezelter 3769
281     \subsection{Force Field Parameters}
282 skuang 3774 For comparison of our new method with previous work, we retain our
283     force field parameters consistent with the results we will compare
284 skuang 3775 with. The Lennard-Jones fluid used here for argon , and reduced unit
285     results are reported for direct comparison purpose.
286 gezelter 3769
287 skuang 3774 As for our water simulations, SPC/E model is used throughout this work
288     for consistency. Previous work for transport properties of SPC/E water
289 skuang 3775 model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
290     that unnecessary repetition of previous methods can be avoided.
291 gezelter 3769
292 skuang 3774 The Au-Au interaction parameters in all simulations are described by
293     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
294     QSC potentials include zero-point quantum corrections and are
295 gezelter 3769 reparametrized for accurate surface energies compared to the
296 skuang 3775 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
297     Spohr potential was adopted\cite{ISI:000167766600035} to depict
298     Au-H$_2$O interactions.
299 gezelter 3769
300 skuang 3774 The small organic molecules included in our simulations are the Au
301     surface capping agent butanethiol and liquid hexane and toluene. The
302     United-Atom
303     models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
304     for these components were used in this work for better computational
305     efficiency, while maintaining good accuracy. We refer readers to our
306     previous work\cite{kuang:AuThl} for further details of these models,
307     as well as the interactions between Au and the above organic molecule
308     components.
309 gezelter 3769
310 skuang 3774 \subsection{Thermal conductivities}
311 skuang 3775 When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
312     impose kinetic energy transfer, the method can be used for thermal
313     conductivity computations. Similar to previous RNEMD methods, we
314     assume linear response of the temperature gradient with respect to the
315     thermal flux in general case. And the thermal conductivity ($\lambda$)
316     can be obtained with the imposed kinetic energy flux and the measured
317     thermal gradient:
318     \begin{equation}
319     J_z = -\lambda \frac{\partial T}{\partial z}
320     \end{equation}
321     Like other imposed-flux methods, the energy flux was calculated using
322     the total non-physical energy transferred (${E_{total}}$) from slab
323     ``c'' to slab ``h'', which is recorded throughout a simulation, and
324     the time for data collection $t$:
325     \begin{equation}
326     J_z = \frac{E_{total}}{2 t L_x L_y}
327     \end{equation}
328     where $L_x$ and $L_y$ denotes the dimensions of the plane in a
329     simulation cell perpendicular to the thermal gradient, and a factor of
330     two in the denominator is present for the heat transport occurs in
331     both $+z$ and $-z$ directions. The temperature gradient
332     ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
333     regression of the temperature profile, which is recorded during a
334     simulation for each slab in a cell. For Lennard-Jones simulations,
335     thermal conductivities are reported in reduced units
336     (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
337    
338 skuang 3774 \subsection{Shear viscosities}
339 skuang 3775 Alternatively, the method can carry out shear viscosity calculations
340     by switching off $J_z$. One can specify the vector
341     $\vec{j}_z(\vec{p})$ by choosing the three components
342     respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
343     set to zero. Although for isotropic systems, the direction of
344     $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, the ability
345     of arbitarily specifying the vector direction in our method provides
346     convenience in anisotropic simulations.
347    
348     Similar to thermal conductivity computations, linear response of the
349     momentum gradient with respect to the shear stress is assumed, and the
350     shear viscosity ($\eta$) can be obtained with the imposed momentum
351     flux (e.g. in $x$ direction) and the measured gradient:
352     \begin{equation}
353     j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
354     \end{equation}
355     where the flux is similarly defined:
356     \begin{equation}
357     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
358     \end{equation}
359     with $P_x$ being the total non-physical momentum transferred within
360     the data collection time. Also, the velocity gradient
361     ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
362     regression of the $x$ component of the mean velocity, $\langle
363     v_x\rangle$, in each of the bins. For Lennard-Jones simulations, shear
364     viscosities are reported in reduced units
365     (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
366    
367 skuang 3774 \subsection{Interfacial friction and Slip length}
368 skuang 3775 While the shear stress results in a velocity gradient within bulk
369     fluid phase, its effect at a solid-liquid interface could vary due to
370     the interaction strength between the two phases. The interfacial
371     friction coefficient $\kappa$ is defined to relate the shear stress
372     (e.g. along $x$-axis) and the relative fluid velocity tangent to the
373     interface:
374     \begin{equation}
375     j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
376     \end{equation}
377     Under ``stick'' boundary condition, $\Delta v_x|_{interface}
378     \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
379     ``slip'' boundary condition at the solid-liquid interface, $\kappa$
380     becomes finite. To characterize the interfacial boundary conditions,
381     slip length ($\delta$) is defined using $\kappa$ and the shear
382     viscocity of liquid phase ($\eta$):
383     \begin{equation}
384     \delta = \frac{\eta}{\kappa}
385     \end{equation}
386     so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
387     and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
388     illustrates how this quantity is defined and computed for a
389 skuang 3776 solid-liquid interface. [MAY INCLUDE A SNAPSHOT IN FIGURE]
390 gezelter 3769
391 skuang 3775 \begin{figure}
392     \includegraphics[width=\linewidth]{defDelta}
393     \caption{The slip length $\delta$ can be obtained from a velocity
394 skuang 3780 profile of a solid-liquid interface simulation. An example of
395     Au/hexane interfaces is shown. Calculation for the left side is
396     illustrated. The right side is similar to the left side.}
397 skuang 3775 \label{slipLength}
398     \end{figure}
399 gezelter 3769
400 skuang 3775 In our method, a shear stress can be applied similar to shear
401     viscosity computations by applying an unphysical momentum flux
402     (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
403     shown in Figure \ref{slipLength}, in which the velocity gradients
404     within liquid phase and velocity difference at the liquid-solid
405     interface can be measured respectively. Further calculations and
406     characterizations of the interface can be carried out using these
407     data.
408    
409 skuang 3776 \section{Results and Discussions}
410     \subsection{Lennard-Jones fluid}
411     Our orthorhombic simulation cell of Lennard-Jones fluid has identical
412     parameters to our previous work\cite{kuang:164101} to facilitate
413     comparison. Thermal conductivitis and shear viscosities were computed
414     with the algorithm applied to the simulations. The results of thermal
415     conductivity are compared with our previous NIVS algorithm. However,
416     since the NIVS algorithm could produce temperature anisotropy for
417     shear viscocity computations, these results are instead compared to
418     the momentum swapping approaches. Table \ref{LJ} lists these
419     calculations with various fluxes in reduced units.
420 skuang 3770
421 skuang 3776 \begin{table*}
422     \begin{minipage}{\linewidth}
423     \begin{center}
424 gezelter 3769
425 skuang 3776 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
426     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
427     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
428     at various momentum fluxes. The new method yields similar
429     results to previous RNEMD methods. All results are reported in
430     reduced unit. Uncertainties are indicated in parentheses.}
431    
432     \begin{tabular}{cccccc}
433     \hline\hline
434     \multicolumn{2}{c}{Momentum Exchange} &
435     \multicolumn{2}{c}{$\lambda^*$} &
436     \multicolumn{2}{c}{$\eta^*$} \\
437     \hline
438 skuang 3779 Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
439 skuang 3776 NIVS & This work & Swapping & This work \\
440     \hline
441     0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
442     0.232 & 0.09 & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
443     0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
444     0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
445     1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
446     \hline\hline
447     \end{tabular}
448     \label{LJ}
449     \end{center}
450     \end{minipage}
451     \end{table*}
452 gezelter 3769
453 skuang 3776 \subsubsection{Thermal conductivity}
454     Our thermal conductivity calculations with this method yields
455     comparable results to the previous NIVS algorithm. This indicates that
456     the thermal gradients rendered using this method are also close to
457     previous RNEMD methods. Simulations with moderately higher thermal
458     fluxes tend to yield more reliable thermal gradients and thus avoid
459     large errors, while overly high thermal fluxes could introduce side
460     effects such as non-linear temperature gradient response or
461     inadvertent phase transitions.
462 gezelter 3769
463 skuang 3776 Since the scaling operation is isotropic in this method, one does not
464     need extra care to ensure temperature isotropy between the $x$, $y$
465     and $z$ axes, while thermal anisotropy might happen if the criteria
466     function for choosing scaling coefficients does not perform as
467     expected. Furthermore, this method avoids inadvertent concomitant
468     momentum flux when only thermal flux is imposed, which could not be
469     achieved with swapping or NIVS approaches. The thermal energy exchange
470 skuang 3778 in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
471 skuang 3776 or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
472     P^\alpha$) would not obtain this result unless thermal flux vanishes
473     (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which are trivial to apply a
474     thermal flux). In this sense, this method contributes to having
475     minimal perturbation to a simulation while imposing thermal flux.
476    
477     \subsubsection{Shear viscosity}
478     Table \ref{LJ} also compares our shear viscosity results with momentum
479     swapping approach. Our calculations show that our method predicted
480     similar values for shear viscosities to the momentum swapping
481     approach, as well as the velocity gradient profiles. Moderately larger
482     momentum fluxes are helpful to reduce the errors of measured velocity
483     gradients and thus the final result. However, it is pointed out that
484     the momentum swapping approach tends to produce nonthermal velocity
485     distributions.\cite{Maginn:2010}
486    
487     To examine that temperature isotropy holds in simulations using our
488     method, we measured the three one-dimensional temperatures in each of
489     the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
490     temperatures were calculated after subtracting the effects from bulk
491     velocities of the slabs. The one-dimensional temperature profiles
492     showed no observable difference between the three dimensions. This
493     ensures that isotropic scaling automatically preserves temperature
494     isotropy and that our method is useful in shear viscosity
495     computations.
496    
497 gezelter 3769 \begin{figure}
498 skuang 3776 \includegraphics[width=\linewidth]{tempXyz}
499 skuang 3777 \caption{Unlike the previous NIVS algorithm, the new method does not
500     produce a thermal anisotropy. No temperature difference between
501     different dimensions were observed beyond the magnitude of the error
502     bars. Note that the two ``hotter'' regions are caused by the shear
503     stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
504     an effect that only observed in our methods.}
505 skuang 3776 \label{tempXyz}
506 gezelter 3769 \end{figure}
507    
508 skuang 3776 Furthermore, the velocity distribution profiles are tested by imposing
509     a large shear stress into the simulations. Figure \ref{vDist}
510     demonstrates how our method is able to maintain thermal velocity
511     distributions against the momentum swapping approach even under large
512     imposed fluxes. Previous swapping methods tend to deplete particles of
513     positive velocities in the negative velocity slab (``c'') and vice
514     versa in slab ``h'', where the distributions leave a notch. This
515     problematic profiles become significant when the imposed-flux becomes
516     larger and diffusions from neighboring slabs could not offset the
517     depletion. Simutaneously, abnormal peaks appear corresponding to
518     excessive velocity swapped from the other slab. This nonthermal
519     distributions limit applications of the swapping approach in shear
520     stress simulations. Our method avoids the above problematic
521     distributions by altering the means of applying momentum
522     fluxes. Comparatively, velocity distributions recorded from
523     simulations with our method is so close to the ideal thermal
524     prediction that no observable difference is shown in Figure
525     \ref{vDist}. Conclusively, our method avoids problems happened in
526     previous RNEMD methods and provides a useful means for shear viscosity
527     computations.
528 gezelter 3769
529 skuang 3776 \begin{figure}
530     \includegraphics[width=\linewidth]{velDist}
531 skuang 3777 \caption{Velocity distributions that develop under the swapping and
532     our methods at high flux. These distributions were obtained from
533 skuang 3779 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
534 skuang 3777 swapping interval of 20 time steps). This is a relatively large flux
535     to demonstrate the nonthermal distributions that develop under the
536     swapping method. Distributions produced by our method are very close
537     to the ideal thermal situations.}
538 skuang 3776 \label{vDist}
539     \end{figure}
540 gezelter 3769
541 skuang 3776 \subsection{Bulk SPC/E water}
542 skuang 3778 Since our method was in good performance of thermal conductivity and
543     shear viscosity computations for simple Lennard-Jones fluid, we extend
544     our applications of these simulations to complex fluid like SPC/E
545     water model. A simulation cell with 1000 molecules was set up in the
546     same manner as in \cite{kuang:164101}. For thermal conductivity
547     simulations, measurements were taken to compare with previous RNEMD
548     methods; for shear viscosity computations, simulations were run under
549     a series of temperatures (with corresponding pressure relaxation using
550     the isobaric-isothermal ensemble[CITE NIVS REF 32]), and results were
551     compared to available data from Equilibrium MD methods[CITATIONS].
552 skuang 3777
553 skuang 3776 \subsubsection{Thermal conductivity}
554 skuang 3778 Table \ref{spceThermal} summarizes our thermal conductivity
555     computations under different temperatures and thermal gradients, in
556     comparison to the previous NIVS results\cite{kuang:164101} and
557     experimental measurements\cite{WagnerKruse}. Note that no appreciable
558     drift of total system energy or temperature was observed when our
559     method is applied, which indicates that our algorithm conserves total
560     energy even for systems involving electrostatic interactions.
561 gezelter 3769
562 skuang 3778 Measurements using our method established similar temperature
563     gradients to the previous NIVS method. Our simulation results are in
564     good agreement with those from previous simulations. And both methods
565     yield values in reasonable agreement with experimental
566     values. Simulations using moderately higher thermal gradient or those
567     with longer gradient axis ($z$) for measurement seem to have better
568     accuracy, from our results.
569    
570     \begin{table*}
571     \begin{minipage}{\linewidth}
572     \begin{center}
573    
574     \caption{Thermal conductivity of SPC/E water under various
575     imposed thermal gradients. Uncertainties are indicated in
576     parentheses.}
577    
578     \begin{tabular}{ccccc}
579     \hline\hline
580     $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
581     {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
582     (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
583     Experiment\cite{WagnerKruse} \\
584     \hline
585     300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\
586     318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
587     & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
588 skuang 3779 & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
589     twice as long.} & & \\
590 skuang 3778 \hline\hline
591     \end{tabular}
592     \label{spceThermal}
593     \end{center}
594     \end{minipage}
595     \end{table*}
596    
597 skuang 3776 \subsubsection{Shear viscosity}
598 skuang 3778 The improvement our method achieves for shear viscosity computations
599     enables us to apply it on SPC/E water models. The series of
600     temperatures under which our shear viscosity calculations were carried
601     out covers the liquid range under normal pressure. Our simulations
602     predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
603 skuang 3779 (Table \ref{spceShear}). Considering subtlties such as temperature or
604     pressure/density errors in these two series of measurements, our
605     results show no significant difference from those with EMD
606     methods. Since each value reported using our method takes only one
607     single trajectory of simulation, instead of average from many
608     trajectories when using EMD, our method provides an effective means
609     for shear viscosity computations.
610 gezelter 3769
611 skuang 3778 \begin{table*}
612     \begin{minipage}{\linewidth}
613     \begin{center}
614    
615     \caption{Computed shear viscosity of SPC/E water under different
616     temperatures. Results are compared to those obtained with EMD
617     method[CITATION]. Uncertainties are indicated in parentheses.}
618    
619     \begin{tabular}{cccc}
620     \hline\hline
621     $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
622     {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
623     (K) & (10$^{10}$s$^{-1}$) & This work & Previous simulations[CITATION]\\
624     \hline
625     273 & & 1.218(0.004) & \\
626     & & 1.140(0.012) & \\
627     303 & & 0.646(0.008) & \\
628     318 & & 0.536(0.007) & \\
629     & & 0.510(0.007) & \\
630     & & & \\
631     333 & & 0.428(0.002) & \\
632     363 & & 0.279(0.014) & \\
633     & & 0.306(0.001) & \\
634     \hline\hline
635     \end{tabular}
636     \label{spceShear}
637     \end{center}
638     \end{minipage}
639     \end{table*}
640 gezelter 3769
641 skuang 3776 [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
642 skuang 3777 [PUT RESULTS AND FIGURE HERE IF IT WORKS]
643 skuang 3779 \subsection{Interfacial frictions and slip lengths}
644     An attractive aspect of our method is the ability to apply momentum
645     and/or thermal flux in nonhomogeneous systems, where molecules of
646     different identities (or phases) are segregated in different
647     regions. We have previously studied the interfacial thermal transport
648     of a series of metal gold-liquid
649     surfaces\cite{kuang:164101,kuang:AuThl}, and attemptions have been
650     made to investigate the relationship between this phenomenon and the
651     interfacial frictions.
652 gezelter 3769
653 skuang 3779 Table \ref{etaKappaDelta} includes these computations and previous
654     calculations of corresponding interfacial thermal conductance. For
655     bare Au(111) surfaces, slip boundary conditions were observed for both
656     organic and aqueous liquid phases, corresponding to previously
657     computed low interfacial thermal conductance. Instead, the butanethiol
658     covered Au(111) surface appeared to be sticky to the organic liquid
659     molecules in our simulations. We have reported conductance enhancement
660     effect for this surface capping agent,\cite{kuang:AuThl} and these
661     observations have a qualitative agreement with the thermal conductance
662     results. This agreement also supports discussions on the relationship
663     between surface wetting and slip effect and thermal conductance of the
664     interface.[CITE BARRAT, GARDE]
665 skuang 3776
666 gezelter 3769 \begin{table*}
667     \begin{minipage}{\linewidth}
668     \begin{center}
669    
670 skuang 3779 \caption{Computed interfacial friction coefficient values for
671     interfaces with various components for liquid and solid
672     phase. Error estimates are indicated in parentheses.}
673 gezelter 3769
674 skuang 3779 \begin{tabular}{llcccccc}
675 gezelter 3769 \hline\hline
676 skuang 3779 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
677     & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
678     \cite{kuang:164101}.} \\
679 skuang 3780 surface & molecules & K & MPa & mPa$\cdot$s & Pa$\cdot$s/m & nm
680     & MW/m$^2$/K \\
681 gezelter 3769 \hline
682 skuang 3779 Au(111) & hexane & 200 & 1.08 & 0.20() & 5.3$\times$10$^4$() &
683     3.7 & 46.5 \\
684     & & & 2.15 & 0.14() & 5.3$\times$10$^4$() &
685     2.7 & \\
686     Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.29() & $\infty$ & 0 &
687     131 \\
688     & & & 5.39 & 0.32() & $\infty$ & 0 &
689     \\
690 gezelter 3769 \hline
691 skuang 3779 Au(111) & toluene & 200 & 1.08 & 0.72() & 1.?$\times$10$^5$() &
692     4.6 & 70.1 \\
693     & & & 2.16 & 0.54() & 1.?$\times$10$^5$() &
694     4.9 & \\
695     Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.98() & $\infty$ & 0
696     & 187 \\
697     & & & 10.8 & 0.99() & $\infty$ & 0
698     & \\
699     \hline
700     Au(111) & water & 300 & 1.08 & 0.40() & 1.9$\times$10$^4$() &
701     20.7 & 1.65 \\
702     & & & 2.16 & 0.79() & 1.9$\times$10$^4$() &
703     41.9 & \\
704     \hline
705     ice(basal) & water & 225 & 19.4 & 15.8() & $\infty$ & 0 & \\
706 gezelter 3769 \hline\hline
707     \end{tabular}
708 skuang 3779 \label{etaKappaDelta}
709 gezelter 3769 \end{center}
710     \end{minipage}
711     \end{table*}
712    
713 skuang 3779 An interesting effect alongside the surface friction change is
714     observed on the shear viscosity of liquids in the regions close to the
715     solid surface. Note that $\eta$ measured near a ``slip'' surface tends
716     to be smaller than that near a ``stick'' surface. This suggests that
717     an interface could affect the dynamic properties on its neighbor
718     regions. It is known that diffusions of solid particles in liquid
719     phase is affected by their surface conditions (stick or slip
720     boundary).[CITE SCHMIDT AND SKINNER] Our observations could provide
721     support to this phenomenon.
722 gezelter 3769
723 skuang 3779 In addition to these previously studied interfaces, we attempt to
724     construct ice-water interfaces and the basal plane of ice lattice was
725     first studied. In contrast to the Au(111)/water interface, where the
726     friction coefficient is relatively small and large slip effect
727     presents, the ice/liquid water interface demonstrates strong
728     interactions and appears to be sticky. The supercooled liquid phase is
729     an order of magnitude viscous than measurements in previous
730     section. It would be of interst to investigate the effect of different
731     ice lattice planes (such as prism surface) on interfacial friction and
732     corresponding liquid viscosity.
733 gezelter 3769
734 skuang 3776 \section{Conclusions}
735 skuang 3779 Our simulations demonstrate the validity of our method in RNEMD
736     computations of thermal conductivity and shear viscosity in atomic and
737     molecular liquids. Our method maintains thermal velocity distributions
738     and avoids thermal anisotropy in previous NIVS shear stress
739     simulations, as well as retains attractive features of previous RNEMD
740     methods. There is no {\it a priori} restrictions to the method to be
741     applied in various ensembles, so prospective applications to
742     extended-system methods are possible.
743 gezelter 3769
744 skuang 3779 Furthermore, using this method, investigations can be carried out to
745     characterize interfacial interactions. Our method is capable of
746     effectively imposing both thermal and momentum flux accross an
747     interface and thus facilitates studies that relates dynamic property
748     measurements to the chemical details of an interface.
749 gezelter 3769
750 skuang 3779 Another attractive feature of our method is the ability of
751     simultaneously imposing thermal and momentum flux in a
752     system. potential researches that might be benefit include complex
753     systems that involve thermal and momentum gradients. For example, the
754     Soret effects under a velocity gradient would be of interest to
755     purification and separation researches.
756 gezelter 3769
757     \section{Acknowledgments}
758     Support for this project was provided by the National Science
759     Foundation under grant CHE-0848243. Computational time was provided by
760     the Center for Research Computing (CRC) at the University of Notre
761     Dame.
762    
763     \newpage
764    
765 skuang 3770 \bibliography{stokes}
766 gezelter 3769
767     \end{doublespace}
768     \end{document}