ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
(Generate patch)

Comparing stokes/stokes.tex (file contents):
Revision 3785 by skuang, Tue Dec 13 23:31:35 2011 UTC vs.
Revision 3786 by gezelter, Wed Dec 14 21:59:12 2011 UTC

# Line 28 | Line 28
28  
29   \begin{document}
30  
31 < \title{A minimal perturbation approach to RNEMD able to simultaneously
32 < impose thermal and momentum gradients}
31 > \title{Velocity Shearing and Scaling RNEMD: a minimally perturbing
32 >  method for producing temperature and momentum gradients}
33  
34   \author{Shenyu Kuang and J. Daniel
35   Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
# Line 46 | Line 46 | Notre Dame, Indiana 46556}
46   \begin{abstract}
47    We present a new method for introducing stable nonequilibrium
48    velocity and temperature gradients in molecular dynamics simulations
49 <  of heterogeneous systems. This method conserves the linear momentum
50 <  and total energy of the system and improves previous Reverse
51 <  Non-Equilibrium Molecular Dynamics (RNEMD) methods while maintaining
52 <  thermal velocity distributions. It also avoid thermal anisotropy
53 <  occured in previous NIVS simulations by using isotropic velocity
54 <  scaling on the molecules in specific regions of a system. To test
55 <  the method, we have computed the thermal conductivity and shear
49 >  of heterogeneous systems. This method conserves both the linear
50 >  momentum and total energy of the system and improves previous
51 >  reverse non-equilibrium molecular dynamics (RNEMD) methods while
52 >  retaining equilibrium thermal velocity distributions in each region
53 >  of the system.  The new method avoids the thermal anisotropy
54 >  produced by previous methods by using isotropic velocity scaling and
55 >  shearing on all of the molecules in specific regions. To test the
56 >  method, we have computed the thermal conductivity and shear
57    viscosity of model liquid systems as well as the interfacial
58 <  frictions of a series of metal/liquid interfaces. Its ability to
59 <  combine the thermal and momentum gradients allows us to obtain shear
60 <  viscosity data for a range of temperatures in only one trajectory.
61 <
58 >  friction coeefficients of a series of solid / liquid interfaces. The
59 >  method's ability to combine the thermal and momentum gradients
60 >  allows us to obtain shear viscosity data for a range of temperatures
61 >  from a single trajectory.
62   \end{abstract}
63  
64   \newpage
# Line 69 | Line 70 | Imposed-flux methods in Molecular Dynamics (MD)
70   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71  
72   \section{Introduction}
73 < Imposed-flux methods in Molecular Dynamics (MD)
74 < simulations\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
75 < can establish steady state systems with an applied flux set vs a
76 < corresponding gradient that can be measured. These methods does not
77 < need many trajectories to provide information of transport properties
78 < of a given system. Thus, they are utilized in computing thermal and
79 < mechanical transfer of homogeneous bulk systems as well as
80 < heterogeneous systems such as solid-liquid
73 > One of the standard ways to compute transport coefficients such as the
74 > viscosities and thermal conductivities of liquids is to use
75 > imposed-flux non-equilibrium molecular dynamics
76 > methods.\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
77 > These methods establish stable non-equilibrium conditions in a
78 > simulation box using an applied momentum or thermal flux between
79 > different regions of the box.  The corresponding temperature or
80 > velocity gradients which develop in response to the applied flux is
81 > related (via linear response theory) to the transport coefficient of
82 > interest.  These methods are quite efficient, in that they do not need
83 > many trajectories to provide information about transport properties. To
84 > date, they have been utilized in computing thermal and mechanical
85 > transfer of both homogeneous liquids as well as heterogeneous systems
86 > such as solid-liquid
87   interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
88  
89 < The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that
90 < satisfy linear momentum and total energy conservation of a system when
91 < imposing fluxes in a simulation. Thus they are compatible with various
92 < ensembles, including the micro-canonical (NVE) ensemble, without the
93 < need of an external thermostat. The original approaches proposed by
94 < M\"{u}ller-Plathe {\it et
89 > The reverse non-equilibrium molecular dynamics (RNEMD) methods utilize
90 > additional constraints that ensure conservation of linear momentum and
91 > total energy of the system while imposing the desired flux.  The RNEMD
92 > methods are therefore capable of sampling various thermodynamically
93 > relevent ensembles, including the micro-canonical (NVE) ensemble,
94 > without resorting to an external thermostat. The original approaches
95 > proposed by M\"{u}ller-Plathe {\it et
96    al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
97 < momentum swapping for generating energy/momentum fluxes, which can
98 < also be compatible with particles of different identities. Although
99 < simple to implement in a simulation, this approach can create
100 < nonthermal velocity distributions, as discovered by Tenney and
101 < Maginn\cite{Maginn:2010}. Furthermore, this approach is less efficient
102 < for kinetic energy transfer between particles of different identities,
103 < especially when the mass difference between the particles becomes
104 < significant. This also limits its applications on heterogeneous
105 < interfacial systems.
97 > momentum swapping moves for generating energy/momentum fluxes.  The
98 > swapping moves can also be made compatible with particles of different
99 > identities. Although the swapping moves are simple to implement in
100 > molecular simulations, Tenney and Maginn have shown that they create
101 > nonthermal velocity distributions.\cite{Maginn:2010} Furthermore, this
102 > approach is not particularly efficient for kinetic energy transfer
103 > between particles of different identities, particularly when the mass
104 > difference between the particles becomes significant.  This problem
105 > makes applying swapping-move RNEMD methods on heterogeneous
106 > interfacial systems somewhat difficult.
107  
108 < Recently, we developed a different approach, using Non-Isotropic
109 < Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose
110 < fluxes. Compared to the momentum swapping move, it scales the velocity
111 < vectors in two separate regions of a simulated system with respective
112 < diagonal scaling matrices. These matrices are determined by solving a
113 < set of equations including linear momentum and kinetic energy
114 < conservation constraints and target flux satisfaction. This method is
115 < able to effectively impose a wide range of kinetic energy fluxes
116 < without obvious perturbation to the velocity distributions of the
117 < simulated systems, regardless of the presence of heterogeneous
118 < interfaces. We have successfully applied this approach in studying the
119 < interfacial thermal conductance at metal-solvent
108 > Recently, we developed a somewhat different approach to applying
109 > thermal fluxes in RNEMD simulation using a Non-Isotropic Velocity
110 > Scaling (NIVS) algorithm.\cite{kuang:164101} This algorithm scales all
111 > atomic velocity vectors in two separate regions of a simulated system
112 > using two diagonal scaling matrices.  The scaling matrices are
113 > determined by solving single quartic equation which includes linear
114 > momentum and kinetic energy conservation constraints as well as the
115 > target thermal flux between the regions.  The NIVS method is able to
116 > effectively impose a wide range of kinetic energy fluxes without
117 > significant perturbation to the velocity distributions away from ideal
118 > Maxwell-Boltzmann distributions, even in the presence of heterogeneous
119 > interfaces. We successfully applied this approach in studying the
120 > interfacial thermal conductance at chemically-capped metal-solvent
121   interfaces.\cite{kuang:AuThl}
122  
123 < However, the NIVS approach has limited applications in imposing
124 < momentum fluxes. Temperature anisotropy could happen under high
125 < momentum fluxes due to the implementation of this algorithm. Thus,
126 < combining thermal and momentum flux is also difficult to obtain with
127 < this approach. However, such combination may provide a means to
128 < simulate thermal/momentum gradient coupled processes such as Soret
129 < effect in liquid flows. Therefore, developing improved approaches to
130 < extend the applications of the imposed-flux method is desirable.
123 > The NIVS approach works very well for preparing stable {\it thermal}
124 > gradients.  However, as we pointed out in the original paper, it has
125 > limited application in imposing {\it linear} momentum fluxes (which
126 > are required for measuring shear viscosities). The reason for this is
127 > that linear momentum flux was being imposed by scaling random
128 > fluctuations of the center of the velocity distributions.  Repeated
129 > application of the original NIVS approach resulted in temperature
130 > anisotropy, i.e. the width of the velocity distributions depended on
131 > coordinates perpendicular to the desired gradient direction.  For this
132 > reason, combining thermal and momentum fluxes was difficult with the
133 > original NIVS algorithm.  However, combinations of thermal and
134 > velocity gradients would provide a means to simulate thermal-linear
135 > coupled processes such as Soret effect in liquid flows. Therefore,
136 > developing improved approaches to the scaling imposed-flux methods
137 > would be useful.
138  
139 < In this paper, we improve the RNEMD methods by proposing a novel
140 < approach to impose fluxes. This approach separate the means of applying
141 < momentum and thermal flux with operations in one time step and thus is
142 < able to simutaneously impose thermal and momentum flux. Furthermore,
143 < the approach retains desirable features of previous RNEMD approaches
144 < and is simpler to implement compared to the NIVS method. In what
145 < follows, we first present the method and its implementation in a
146 < simulation. Then we compare the method on bulk fluids to previous
147 < methods. Also, interfacial frictions are computed for a series of
148 < interfaces.
139 > In this paper, we improve the RNEMD methods by introducing a novel
140 > approach to creating imposed fluxes. This approach separates the
141 > shearing and scaling of the velocity distributions in different
142 > spatial regions and can apply both transformations within a single
143 > time step.  The approach retains desirable features of previous RNEMD
144 > approaches and is simpler to implement compared to the earlier NIVS
145 > method. In what follows, we first present the Shearing-and-Scaling
146 > (SS) RNEMD method and its implementation in a simulation. Then we
147 > compare the SS-RNEMD method in bulk fluids to previous methods.  We
148 > also compute interfacial frictions are computed for a series of
149 > heterogeneous interfaces.
150  
151   \section{Methodology}
152 < Similar to the NIVS method,\cite{kuang:164101} we consider a
153 < periodic system divided into a series of slabs along a certain axis
154 < (e.g. $z$). The unphysical thermal and/or momentum flux is designated
155 < from the center slab to one of the end slabs, and thus the thermal
156 < flux results in a lower temperature of the center slab than the end
157 < slab, and the momentum flux results in negative center slab momentum
158 < with positive end slab momentum (unless these fluxes are set
159 < negative). Therefore, the center slab is denoted as ``$c$'', while the
160 < end slab as ``$h$''.
152 > In an approach similar to the earlier NIVS method,\cite{kuang:164101}
153 > we consider a periodic system which has been divided into a series of
154 > slabs along a single axis (e.g. $z$). The unphysical thermal and
155 > momentum fluxes are applied from one of the end slabs to the center
156 > slab, and thus the thermal flux produces a higher temperature in the
157 > center slab than in the end slab, and the momentum flux results in a
158 > center slab moving along the positive $y$ axis (see Fig. \ref{cartoon}).
159 > The applied fluxes can be set to negative values if the reverse
160 > gradients are desired.  For convenience the center slab is denoted as
161 > the {\it hot} or {\it ``H''} slab, while the end slab is denoted {\it
162 >  ``C''} (or {\it cold}).
163  
164 < To impose these fluxes, we periodically apply different set of
165 < operations on velocities of particles {$i$} within the center slab and
166 < those of particles {$j$} within the end slab:
164 > \begin{figure}
165 > \includegraphics[width=\linewidth]{cartoon}
166 > \caption{The SS-RNEMD approach we are introducing imposes unphysical
167 >  transfer of both momentum and kinetic energy between a ``hot'' slab
168 >  and a ``cold'' slab in the simulation box.  The molecular system
169 >  responds to this imposed flux by generating both momentum and
170 >  temperature gradients.  The slope of the gradients can then be used
171 >  to compute transport properties (e.g. shear viscosity and thermal
172 >  conductivity).}
173 > \label{cartoon}
174 > \end{figure}
175 >
176 > To impose these fluxes, we periodically apply a set of operations on
177 > the velocities of particles {$i$} within the cold slab and a separate
178 > operation on particles {$j$} within the hot slab.
179   \begin{eqnarray}
180   \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
181    \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
182   \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
183    \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
184   \end{eqnarray}
185 < where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes
186 < the instantaneous bulk velocity of slabs $c$ and $h$ respectively
187 < before an operation is applied. When a momentum flux $\vec{j}_z(\vec{p})$
188 < presents, these bulk velocities would have a corresponding change
189 < ($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's
190 < second law:
185 > where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denote
186 > the instantaneous average velocity of the molecules within slabs $C$
187 > and $H$ respectively.  When a momentum flux $\vec{j}_z(\vec{p})$ is
188 > present, these slab-averaged velocities also get corresponding
189 > incremental changes ($\vec{a}_c$ and $\vec{a}_h$ respectively) that
190 > are applied to all particles within each slab.  The incremental
191 > changes are obtained using Newton's second law:
192   \begin{eqnarray}
193 < M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\
193 > M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \label{eq:newton1} \\
194   M_h \vec{a}_h & = &  \vec{j}_z(\vec{p}) \Delta t
195 + \label{eq:newton2}
196   \end{eqnarray}
197   where $M$ denotes total mass of particles within a slab:
198   \begin{eqnarray}
# Line 167 | Line 201 | The above operations already conserve the linear momen
201   \end{eqnarray}
202   and $\Delta t$ is the interval between two separate operations.
203  
204 < The above operations already conserve the linear momentum of a
205 < periodic system. To further satisfy total energy conservation as well
206 < as to impose the thermal flux $J_z$, the following equations are
207 < included as well:
208 < [MAY PUT EXTRA MATH IN SUPPORT INFO OR APPENDIX]
204 > The operations in Eqs. \ref{eq:newton1} and \ref{eq:newton2} already
205 > conserve the linear momentum of a periodic system.  To further satisfy
206 > total energy conservation as well as to impose the thermal flux $J_z$,
207 > the following constraint equations must be solved for the two scaling
208 > variables $c$ and $h$:
209   \begin{eqnarray}
210   K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
211   \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
212   K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
213 < \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2
213 > \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2.
214 > \label{constraint}
215   \end{eqnarray}
216 < where $K_c$ and $K_h$ denotes translational kinetic energy of slabs
217 < $c$ and $h$ respectively before an operation is applied. These
218 < translational kinetic energy conservation equations are sufficient to
219 < ensure total energy conservation, as the operations applied in our
220 < method do not change the kinetic energy related to other degrees of
221 < freedom or the potential energy of a system, given that its potential
187 < energy does not depend on particle velocity.
216 > Here $K_c$ and $K_h$ denote the translational kinetic energy of slabs
217 > $C$ and $H$ respectively. These conservation equations are sufficient
218 > to ensure total energy conservation, as the operations applied in our
219 > method do not change the kinetic energy related to orientational
220 > degrees of freedom or the potential energy of a system (as long as the
221 > potential energy is independent of particle velocity).
222  
223 < The above sets of equations are sufficient to determine the velocity
224 < scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
225 < $\vec{a}_h$. Note that there are two roots respectively for $c$ and
226 < $h$. However, the positive roots (which are closer to 1) are chosen so
227 < that the perturbations to a system can be reduced to a minimum. Figure
228 < \ref{method} illustrates the implementation sketch of this algorithm
229 < in an individual step.
223 > Equations \ref{eq:newton1}-\ref{constraint} are sufficient to
224 > determine the velocity scaling coefficients ($c$ and $h$) as well as
225 > $\vec{a}_c$ and $\vec{a}_h$. Note that there are usually two roots
226 > respectively for $c$ and $h$. However, the positive roots (which are
227 > closer to 1) are chosen so that the perturbations to the system are
228 > minimal.  Figure \ref{method} illustrates the implementation of this
229 > algorithm in an individual step.
230  
231   \begin{figure}
232 < \includegraphics[width=\linewidth]{method}
233 < \caption{Illustration of the implementation of the algorithm in a
234 <  single step. Starting from an ideal velocity distribution, the
235 <  transformation is used to apply the effect of both a thermal and a
236 <  momentum flux from the ``c'' slab to the ``h'' slab. As the figure
237 <  shows, thermal distributions can preserve after this operation.}
232 > \centering
233 > \includegraphics[width=5in]{method}
234 > \caption{Illustration of a single step implementation of the
235 >  algorithm.  Starting with the velocity distributions for the two
236 >  slabs in a shearing fluid, the transformation is used to apply the
237 >  effect of both a thermal and a momentum flux from the ``c'' slab to
238 >  the ``h'' slab. As the figure shows, gaussian distributions are
239 >  preserved by both the scaling and shearing operations.}
240   \label{method}
241   \end{figure}
242  
243 < By implementing these operations at a certain frequency, a steady
244 < thermal and/or momentum flux can be applied and the corresponding
245 < temperature and/or momentum gradients can be established.
243 > By implementing these operations at a fixed frequency, stable thermal
244 > and momentum fluxes can both be applied and the corresponding
245 > temperature and momentum gradients can be established.
246  
247 < Compared to the previous NIVS method, this approach is computationally
248 < more efficient in that only quadratic equations are involved to
249 < determine a set of scaling coefficients, while the NIVS method needs
250 < to solve quartic equations. Furthermore, this method implements
251 < isotropic scaling of velocities in respective slabs, unlike the NIVS,
252 < where an extra criteria function is necessary to choose a set of
253 < coefficients that performs a scaling as isotropic as possible. More
254 < importantly, separating the means of momentum flux imposing from
255 < velocity scaling avoids the underlying cause to thermal anisotropy in
256 < NIVS when applying a momentum flux. And later sections will
257 < demonstrate that this can improve the performance in shear viscosity
222 < simulations.
247 > Compared to the previous NIVS method, the SS-RNEMD approach is
248 > computationally simpler in that only quadratic equations are involved
249 > to determine a set of scaling coefficients, while the NIVS method
250 > required solution of quartic equations. Furthermore, this method
251 > implements {\it isotropic} scaling of velocities in respective slabs,
252 > unlike NIVS, which required extra flexibility in the choice of scaling
253 > coefficients to allow for the energy and momentum constraints.  Most
254 > importantly, separating the linear momentum flux from velocity scaling
255 > avoids the underlying cause of the thermal anisotropy in NIVS.  In
256 > later sections, we demonstrate that this can improve the calculated
257 > shear viscosities from RNEMD simulations.
258  
259 < This approach is advantageous over the original momentum swapping in
260 < many aspects. In one swapping, the velocity vectors involved are
261 < usually very different (or the generated flux is trivial to obtain
262 < gradients), thus the swapping tends to incur perturbations to the
263 < neighbors of the particles involved. Comparatively, our approach
264 < disperse the flux to every selected particle in a slab so that
265 < perturbations in the flux generating region could be
266 < minimized. Additionally, because the momentum swapping steps tend to
267 < result in a nonthermal distribution, when an imposed flux is
268 < relatively large and diffusions from the neighboring slabs could no
269 < longer remedy this effect, problematic distributions would be
270 < observed. In comparison, the operations of our approach has the nature
271 < of preserving the equilibrium velocity distributions (commonly
272 < Maxwell-Boltzmann), and results in later section will illustrate that
273 < this is helpful to retain thermal distributions in a simulation.
259 > The SS-RNEMD approach has advantages over the original momentum
260 > swapping in many respects. In the swapping method, the velocity
261 > vectors involved are usually very different (or the generated flux
262 > would be quite small), thus the swapping tends to create strong
263 > perturbations in the neighborhood of the particles involved. In
264 > comparison, the SS approach distributes the flux widely to all
265 > particles in a slab so that perturbations in the flux generating
266 > region are minimized. Additionally, momentum swapping steps tend to
267 > result in nonthermal velocity distributions when the imposed flux is
268 > large and diffusion from the neighboring slabs cannot carry momentum
269 > away quickly enough.  In comparison, the scaling and shearing moves
270 > made in the SS approach preserve the shapes of the equilibrium
271 > velocity disributions (e.g. Maxwell-Boltzmann).  The results presented
272 > in later sections will illustrate that this is quite helpful in
273 > retaining reasonable thermal distributions in a simulation.
274  
275   \section{Computational Details}
276   The algorithm has been implemented in our MD simulation code,
277 < OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with
278 < previous RNEMD methods or equilibrium MD (EMD) methods in homogeneous fluids
279 < (Lennard-Jones and SPC/E water). And taking advantage of the method,
280 < we simulate the interfacial friction of different heterogeneous
281 < interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid
282 < water).
277 > OpenMD.\cite{Meineke:2005gd,openmd} We will first compare the method
278 > with previous RNEMD methods and equilibrium MD in homogeneous
279 > fluids (Lennard-Jones and SPC/E water).  We have also used the new
280 > method to simulate the interfacial friction of different heterogeneous
281 > interfaces (Au (111) with organic solvents, Au(111) with SPC/E water,
282 > and the Ice Ih - liquid water interface).
283  
284   \subsection{Simulation Protocols}
285 < The systems to be investigated are set up in orthorhombic simulation
285 > The systems we investigated were set up in orthorhombic simulation
286   cells with periodic boundary conditions in all three dimensions. The
287 < $z$ axis of these cells were longer and set as the temperature and/or
288 < momentum gradient axis. And the cells were evenly divided into $N$
289 < slabs along this axis, with various $N$ depending on individual
290 < system. The $x$ and $y$ axis were of the same length in homogeneous
291 < systems or had length scale close to each other where heterogeneous
292 < interfaces presents. In all cases, before introducing a nonequilibrium
293 < method to establish steady thermal and/or momentum gradients for
294 < further measurements and calculations, canonical ensemble with a
295 < Nos\'e-Hoover thermostat\cite{hoover85} and microcanonical ensemble
296 < equilibrations were used before data collections. For SPC/E water
297 < simulations, isobaric-isothermal equilibrations\cite{melchionna93} are
298 < performed before the above to reach normal pressure (1 bar); for
299 < interfacial systems, similar equilibrations are used to relax the
300 < surface tensions of the $xy$ plane.
287 > $z$-axes of these cells were typically quite long and served as the
288 > temperature and/or momentum gradient axes.  The cells were evenly
289 > divided into $N$ slabs along this axis, with $N$ varying for the
290 > individual system. The $x$ and $y$ axes were of similar lengths in all
291 > simulations. In all cases, before introducing a nonequilibrium method
292 > to establish steady thermal and/or momentum gradients, equilibration
293 > simulations were run under the canonical ensemble with a Nos\'e-Hoover
294 > thermostat\cite{hoover85} followed by further equilibration using
295 > standard constant energy (NVE) conditions.  For SPC/E water
296 > simulations, isobaric-isothermal equilibrations\cite{melchionna93}
297 > were performed before equilibration to reach standard densities at
298 > atmospheric pressure (1 bar); for interfacial systems, similar
299 > equilibrations with anisotropic box relaxations are used to relax the
300 > surface tension in the $xy$ plane.
301  
302 < While homogeneous fluid systems can be set up with rather random
303 < configurations, our interfacial systems needs a series of steps to
304 < ensure the interfaces be established properly for computations. The
305 < preparation and equilibration of butanethiol covered gold (111)
306 < surface and further solvation and equilibration process is described
307 < in details as in reference \cite{kuang:AuThl}.
302 > While homogeneous fluid systems can be set up with random
303 > configurations, interfacial systems are typically prepared with a
304 > single crystal face presented perpendicular to the $z$-axis. This
305 > crystal face is aligned in the x-y plane of the periodic cell, and the
306 > solvent occupies the region directly above and below a crystalline
307 > slab.  The preparation and equilibration of butanethiol covered
308 > Au(111) surfaces, as well as the solvation and equilibration processes
309 > used for these interfaces are described in detail in reference
310 > \cite{kuang:AuThl}.
311  
312 < As for the ice/liquid water interfaces, the basal surface of ice
313 < lattice was first constructed. Hirsch {\it et
314 <  al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice
315 < lattices with all possible proton order configurations. We refer to
316 < their results and choose the configuration of the lowest energy after
317 < geometry optimization as the unit cell for our ice lattices. Although
312 > For the ice / liquid water interfaces, the basal surface of ice
313 > lattice was first constructed. Hirsch and Ojam\"{a}e
314 > \cite{doi:10.1021/jp048434u} explored the energetics of ice lattices
315 > with all possible proton ordered configurations. We utilized Hirsch
316 > and Ojam\"{a}e's structure 6 ($P2_12_12_1$) which is an orthorhombic
317 > cell giving a proton-ordered version of Ice Ih.  The basal face of ice
318 > in this unit cell geometry is the $\{0~0~1\}$ face.  Although
319   experimental solid/liquid coexistant temperature under normal pressure
320 < should be close to 273K, Bryk and Haymet's simulations of ice/liquid
321 < water interfaces with different models suggest that for SPC/E, the
322 < most stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
323 < Therefore, our ice/liquid water simulations were carried out at
324 < 225K. To have extra protection of the ice lattice during initial
325 < equilibration (when the randomly generated liquid phase configuration
326 < could release large amount of energy in relaxation), restraints were
327 < applied to the ice lattice to avoid inadvertent melting by the heat
289 < dissipated from the high enery configurations.
290 < [MAY ADD A SNAPSHOT FOR BASAL PLANE]
320 > are close to 273K, Bryk and Haymet's simulations of ice/liquid water
321 > interfaces with different models suggest that for SPC/E, the most
322 > stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
323 > Therefore, our ice/liquid water simulations were carried out at an
324 > average temperature of 225K.  Molecular translation and orientational
325 > restraints were applied in the early stages of equilibration to
326 > prevent melting of the ice slab.  These restraints were removed during
327 > NVT equilibration, well before data collection was carried out.
328  
329   \subsection{Force Field Parameters}
330 < For comparison of our new method with previous work, we retain our
330 > For comparing the SS-RNEMD method with previous work, we retained
331   force field parameters consistent with previous simulations. Argon is
332   the Lennard-Jones fluid used here, and its results are reported in
333 < reduced unit for direct comparison purpose.
333 > reduced units for purposes of direct comparison with previous
334 > calculations.
335  
336 < As for our water simulations, SPC/E model is used throughout this work
337 < for consistency. Previous work for transport properties of SPC/E water
338 < model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
339 < that unnecessary repetition of previous methods can be avoided.
336 > For our water simulations, we utilized the SPC/E model throughout this
337 > work.  Previous work for transport properties of SPC/E water model is
338 > available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct
339 > comparison with previous calculation methods is possible.
340  
341   The Au-Au interaction parameters in all simulations are described by
342   the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
# Line 310 | Line 348 | butanethiol and liquid hexane and toluene. The United-
348  
349   For our gold/organic liquid interfaces, the small organic molecules
350   included in our simulations are the Au surface capping agent
351 < butanethiol and liquid hexane and toluene. The United-Atom
351 > butanethiol as well as liquid hexane and liquid toluene. The
352 > United-Atom
353   models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
354 < for these components were used in this work for better computational
354 > for these components were used in this work for computational
355   efficiency, while maintaining good accuracy. We refer readers to our
356   previous work\cite{kuang:AuThl} for further details of these models,
357   as well as the interactions between Au and the above organic molecule
358   components.
359  
360   \subsection{Thermal conductivities}
361 < When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
362 < impose kinetic energy transfer, the method can be used for thermal
363 < conductivity computations. Similar to previous RNEMD methods, we
364 < assume linear response of the temperature gradient with respect to the
365 < thermal flux in general case. And the thermal conductivity ($\lambda$)
366 < can be obtained with the imposed kinetic energy flux and the measured
361 > When the linear momentum flux $\vec{j}_z(\vec{p})$ is set to zero and
362 > the target $J_z$ is non-zero, SS-RNEMD imposes kinetic energy transfer
363 > between the slabs, which can be used for computation of thermal
364 > conductivities. Similar to previous RNEMD methods, we assume that we
365 > are in the linear response regime of the temperature gradient with
366 > respect to the thermal flux.  The thermal conductivity ($\lambda$) can
367 > be calculated using the imposed kinetic energy flux and the measured
368   thermal gradient:
369   \begin{equation}
370   J_z = -\lambda \frac{\partial T}{\partial z}
371   \end{equation}
372   Like other imposed-flux methods, the energy flux was calculated using
373   the total non-physical energy transferred (${E_{total}}$) from slab
374 < ``c'' to slab ``h'', which is recorded throughout a simulation, and
375 < the time for data collection $t$:
374 > ``c'' to slab ``h'', which was recorded throughout the simulation, and
375 > the total data collection time $t$:
376   \begin{equation}
377 < J_z = \frac{E_{total}}{2 t L_x L_y}
377 > J_z = \frac{E_{total}}{2 t L_x L_y}.
378   \end{equation}
379 < where $L_x$ and $L_y$ denotes the dimensions of the plane in a
379 > Here, $L_x$ and $L_y$ denote the dimensions of the plane in a
380   simulation cell perpendicular to the thermal gradient, and a factor of
381 < two in the denominator is necessary for the heat transport occurs in
382 < both $+z$ and $-z$ directions. The average temperature gradient
381 > two in the denominator is necessary as the heat transport occurs in
382 > both the $+z$ and $-z$ directions. The average temperature gradient
383   ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
384   regression of the temperature profile, which is recorded during a
385   simulation for each slab in a cell. For Lennard-Jones simulations,
# Line 347 | Line 387 | Alternatively, the method can carry out shear viscosit
387   (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
388  
389   \subsection{Shear viscosities}
390 < Alternatively, the method can carry out shear viscosity calculations
391 < by specify a momentum flux. In our algorithm, one can specify the
392 < three components of the flux vector $\vec{j}_z(\vec{p})$
393 < respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
394 < set to zero. For isotropic systems, the direction of
395 < $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
356 < ability of arbitarily specifying the vector direction in our method
357 < could provide convenience in anisotropic simulations.
390 > Alternatively, when the linear momentum flux $\vec{j}_z(\vec{p})$ is
391 > non-zero in either the $x$ or $y$ directions, the method can be used
392 > to compute the shear viscosity.  For isotropic systems, the direction
393 > of $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
394 > ability to arbitarily specify the vector direction in our method could
395 > provide convenience when working with anisotropic interfaces.
396  
397 < Similar to thermal conductivity computations, for a homogeneous
398 < system, linear response of the momentum gradient with respect to the
399 < shear stress is assumed, and the shear viscosity ($\eta$) can be
400 < obtained with the imposed momentum flux (e.g. in $x$ direction) and
401 < the measured gradient:
397 > In a manner similar to the thermal conductivity calculations, a linear
398 > response of the momentum gradient with respect to the shear stress is
399 > assumed, and the shear viscosity ($\eta$) can be obtained with the
400 > imposed momentum flux (e.g. in $x$ direction) and the measured
401 > velocity gradient:
402   \begin{equation}
403   j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
404   \end{equation}
# Line 376 | Line 414 | Although $J_z$ may be switched off for shear viscosity
414   viscosities are also reported in reduced units
415   (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
416  
417 < Although $J_z$ may be switched off for shear viscosity simulations at
418 < a certain temperature, our method's ability to impose both a thermal
419 < and a momentum flux in one simulation allows the combination of a
420 < temperature and a velocity gradient. In this case, since viscosity is
421 < generally a function of temperature, the local viscosity also depends
422 < on the local temperature. Therefore, in one such simulation, viscosity
423 < at $z$ (corresponding to a certain $T$) can be computed with the
424 < applied shear flux and the local velocity gradient (which can be
425 < obtained by finite difference approximation). As a whole, the
426 < viscosity can be mapped out as the function of temperature in one
427 < single trajectory of simulation. Results for shear viscosity
428 < computations of SPC/E water will demonstrate its effectiveness in
429 < detail.
417 > Although $J_z$ may be switched off for shear viscosity simulations,
418 > the SS-RNEMD method allows the user the ability to simultaneously
419 > impose both a thermal and a momentum flux during a single
420 > simulation. This can create system with coincident temperature and a
421 > velocity gradients.  Since the viscosity is generally a function of
422 > temperature, the local viscosity depends on the local temperature in
423 > the fluid. Therefore, in a single simulation, viscosity at $z$
424 > (corresponding to a certain $T$) can be computed with the applied
425 > shear flux and the local velocity gradient (which can be obtained by
426 > finite difference approximation). This means that the temperature
427 > dependence of the viscosity can be mapped out in only one
428 > trajectory. Results for shear viscosity computations of SPC/E water
429 > will demonstrate SS-RNEMD's efficiency in this respect.
430  
431 < \subsection{Interfacial friction and Slip length}
432 < While the shear stress results in a velocity gradient within bulk
433 < fluid phase, its effect at a solid-liquid interface could vary due to
434 < the interaction strength between the two phases. The interfacial
435 < friction coefficient $\kappa$ is defined to relate the shear stress
436 < (e.g. along $x$-axis) with the relative fluid velocity tangent to the
399 < interface:
431 > \subsection{Interfacial friction and slip length}
432 > Shear stress creates a velocity gradient within bulk fluid phases, but
433 > at solid-liquid interfaces, the effects of a shear stress depend on
434 > the molecular details of the interface. The interfacial friction
435 > coefficient, $\kappa$, relates the shear stress (e.g. along the
436 > $x$-axis) with the relative fluid velocity tangent to the interface:
437   \begin{equation}
438 < j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
438 > j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right]
439   \end{equation}
440 < Under ``stick'' boundary condition, $\Delta v_x|_{interface}
441 < \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
442 < ``slip'' boundary conditions at the solid-liquid interfaces, $\kappa$
443 < becomes finite. To characterize the interfacial boundary conditions,
444 < slip length ($\delta$) is defined using $\kappa$ and the shear
445 < viscocity of liquid phase ($\eta$):
440 > where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured
441 > directly adjacent to the interface.  Under ``stick'' boundary
442 > condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads
443 > to $\kappa\rightarrow\infty$. However, for ``slip'' boundary
444 > conditions at a solid-liquid interface, $\kappa$ becomes finite. To
445 > characterize the interfacial boundary conditions, the slip length,
446 > $\delta$, is defined by the ratio of the fluid-phase viscosity to the
447 > friction coefficient of the interface:
448   \begin{equation}
449   \delta = \frac{\eta}{\kappa}
450   \end{equation}
451 < so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
452 < and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
453 < illustrates how this quantity is defined and computed for a
454 < solid-liquid interface. [MAY INCLUDE SNAPSHOT IN FIGURE]
451 > In ``no-slip'' or ``stick'' boundary conditions, $\delta\rightarrow
452 > 0$, and $\delta$ is a measure of how ``slippery'' an interface is.
453 > Figure \ref{slipLength} illustrates how this quantity is defined and
454 > computed for a solid-liquid interface.
455  
456   \begin{figure}
457   \includegraphics[width=\linewidth]{defDelta}
458   \caption{The slip length $\delta$ can be obtained from a velocity
459    profile of a solid-liquid interface simulation, when a momentum flux
460 <  is applied. An example of Au/hexane interfaces is shown, and the
461 <  calculation for the left side is illustrated. The calculation for
462 <  the right side is similar to the left.}
460 >  is applied. The data shown is for a simulated Au/hexane interface.
461 >  The Au crystalline region is moving as a block (lower dots), while
462 >  the measured velocity gradient in the hexane phase is discontinuous
463 >  a the interface.}
464   \label{slipLength}
465   \end{figure}
466  
467 < In our method, a shear stress can be applied similar to shear
468 < viscosity computations by applying an unphysical momentum flux
469 < (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
470 < shown in Figure \ref{slipLength}, in which the velocity gradients
471 < within liquid phase and velocity difference at the liquid-solid
472 < interface can be measured respectively. Further calculations and
473 < characterizations of the interface can be carried out using these
474 < data.
467 > Since the method can be applied for interfaces as well as for bulk
468 > materials, the shear stress is applied in an identical manner to the
469 > shear viscosity computations, e.g. by applying an unphysical momentum
470 > flux, $j_z(\vec{p})$.  With the correct choice of $\vec{p}$ in the
471 > $x-y$ plane, one can compute friction coefficients and slip lengths
472 > for a number of different dragging vectors on a given slab.  The
473 > corresponding velocity profiles can be obtained as shown in Figure
474 > \ref{slipLength}, in which the velocity gradients within the liquid
475 > phase and the velocity difference at the liquid-solid interface can be
476 > easily measured from saved simulation data.
477  
478   \section{Results and Discussions}
479   \subsection{Lennard-Jones fluid}
480 < Our orthorhombic simulation cell of Lennard-Jones fluid has identical
481 < parameters to our previous work\cite{kuang:164101} to facilitate
482 < comparison. Thermal conductivitis and shear viscosities were computed
483 < with the algorithm applied to the simulations. The results of thermal
484 < conductivity are compared with our previous NIVS algorithm. However,
485 < since the NIVS algorithm could produce temperature anisotropy for
486 < shear viscocity computations, these results are instead compared to
487 < the momentum swapping approaches. Table \ref{LJ} lists these
488 < calculations with various fluxes in reduced units.
480 > Our orthorhombic simulation cell for the Lennard-Jones fluid has
481 > identical parameters to our previous work\cite{kuang:164101} to
482 > facilitate comparison. Thermal conductivities and shear viscosities
483 > were computed with the new algorithm applied to the simulations. The
484 > results of thermal conductivity are compared with our previous NIVS
485 > algorithm. However, since the NIVS algorithm produced temperature
486 > anisotropy for shear viscocity computations, these results are instead
487 > compared to the previous momentum swapping approaches. Table \ref{LJ}
488 > lists these values with various fluxes in reduced units.
489  
490   \begin{table*}
491    \begin{minipage}{\linewidth}
# Line 462 | Line 504 | calculations with various fluxes in reduced units.
504          \multicolumn{2}{c}{$\lambda^*$} &
505          \multicolumn{2}{c}{$\eta^*$} \\
506          \hline
507 <        Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
507 >        Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
508          NIVS\cite{kuang:164101} & This work & Swapping & This work \\
509          \hline
510          0.116 & 0.16  & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
# Line 498 | Line 540 | applying a thermal flux). In this sense, this method a
540   or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
541   P^\alpha$) would not achieve this effect unless thermal flux vanishes
542   (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which do not contribute to
543 < applying a thermal flux). In this sense, this method aids to achieve
543 > applying a thermal flux). In this sense, this method aids in achieving
544   minimal perturbation to a simulation while imposing a thermal flux.
545  
546   \subsubsection{Shear viscosity}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines