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

Comparing trunk/nivsRnemd/nivsRnemd.tex (file contents):
Revision 3594 by gezelter, Mon Apr 19 16:36:45 2010 UTC vs.
Revision 3600 by gezelter, Thu Apr 22 20:58:46 2010 UTC

# Line 84 | Line 84 | RNEMD has been widely used to provide computational es
84   j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85   J_z & = & \lambda \frac{\partial T}{\partial z}
86   \end{eqnarray}
87 < RNEMD has been widely used to provide computational estimates of thermal
88 < conductivities and shear viscosities in a wide range of materials,
89 < from liquid copper to monatomic liquids to molecular fluids
90 < (e.g. ionic liquids).\cite{Bedrov:2000-1,Bedrov:2000,Muller-Plathe:2002,ISI:000184808400018,ISI:000231042800044,Maginn:2007,Muller-Plathe:2008,ISI:000258460400020,ISI:000258840700015,ISI:000261835100054}
87 > RNEMD has been widely used to provide computational estimates of
88 > thermal conductivities and shear viscosities in a wide range of
89 > materials, from liquid copper to both monatomic and molecular fluids
90 > (e.g.  ionic
91 > liquids).\cite{Bedrov:2000-1,Bedrov:2000,Muller-Plathe:2002,ISI:000184808400018,ISI:000231042800044,Maginn:2007,Muller-Plathe:2008,ISI:000258460400020,ISI:000258840700015,ISI:000261835100054}
92  
93   \begin{figure}
94   \includegraphics[width=\linewidth]{thermalDemo}
# Line 103 | Line 104 | stress) and it is typically much easier to compute mom
104   RNEMD is preferable in many ways to the forward NEMD
105   methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
106   because it imposes what is typically difficult to measure (a flux or
107 < stress) and it is typically much easier to compute momentum gradients
108 < or strains (the response).  For similar reasons, RNEMD is also
107 > stress) and it is typically much easier to compute the response
108 > (momentum gradients or strains.  For similar reasons, RNEMD is also
109   preferable to slowly-converging equilibrium methods for measuring
110   thermal conductivity and shear viscosity (using Green-Kubo
111   relations\cite{daivis:541,mondello:9327} or the Helfand moment
# Line 128 | Line 129 | scaling (NIVS-RNEMD) which retains the desirable featu
129   and self-adjusting metrics for retaining the usability of the method.
130  
131   In this paper, we develop and test a method for non-isotropic velocity
132 < scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
132 > scaling (NIVS) which retains the desirable features of RNEMD
133   (conservation of linear momentum and total energy, compatibility with
134   periodic boundary conditions) while establishing true thermal
135 < distributions in each of the two slabs.  In the next section, we
135 > distributions in each of the two slabs. In the next section, we
136   present the method for determining the scaling constraints.  We then
137 < test the method on both single component, multi-component, and
138 < non-isotropic mixtures and show that it is capable of providing
137 > test the method on both liquids and solids as well as a non-isotropic
138 > liquid-solid interface and show that it is capable of providing
139   reasonable estimates of the thermal conductivity and shear viscosity
140 < in these cases.
140 > in all of these cases.
141  
142   \section{Methodology}
143   We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
# Line 156 | Line 157 | where ${x, y, z}$ are a set of 3 scaling variables for
157   0 & 0 & z \\
158   \end{array} \right) \cdot \vec{v}_i
159   \end{equation}
160 < where ${x, y, z}$ are a set of 3 scaling variables for each of the
161 < three directions in the system.  Likewise, the molecules $\{j\}$
162 < located in the hot slab will see a concomitant scaling of velocities,
160 > where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
161 > of the three directions in the system.  Likewise, the molecules
162 > $\{j\}$ located in the hot slab will see a concomitant scaling of
163 > velocities,
164   \begin{equation}
165   \vec{v}_j \leftarrow \left( \begin{array}{ccc}
166   x^\prime & 0 & 0 \\
# Line 204 | Line 206 | Substituting in the expressions for the hot scaling pa
206   ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
207   {\it constraint ellipsoid}:
208   \begin{equation}
209 < \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
209 > \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
210 >  c_\alpha \right) = 0
211   \label{eq:constraintEllipsoid}
212   \end{equation}
213   where the constants are obtained from the instantaneous values of the
# Line 217 | Line 220 | cold slab scaling parameters which can be applied whil
220   \label{eq:constraintEllipsoidConsts}
221   \end{eqnarray}
222   This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
223 < cold slab scaling parameters which can be applied while preserving
224 < both linear momentum in all three directions as well as total kinetic
225 < energy.
223 > cold slab scaling parameters which, when applied, preserve the linear
224 > momentum of the system in all three directions as well as total
225 > kinetic energy.
226  
227 < The goal of using velocity scaling variables is to transfer linear
228 < momentum or kinetic energy from the cold slab to the hot slab.  If the
229 < hot and cold slabs are separated along the z-axis, the energy flux is
230 < given simply by the decrease in kinetic energy of the cold bin:
227 > The goal of using these velocity scaling variables is to transfer
228 > linear momentum or kinetic energy from the cold slab to the hot slab.
229 > If the hot and cold slabs are separated along the z-axis, the energy
230 > flux is given simply by the decrease in kinetic energy of the cold
231 > bin:
232   \begin{equation}
233   (1-x^2) K_c^x  + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
234   \end{equation}
235   The expression for the energy flux can be re-written as another
236   ellipsoid centered on $(x,y,z) = 0$:
237   \begin{equation}
238 < x^2 K_c^x + y^2 K_c^y + z^2 K_c^z = K_c^x + K_c^y + K_c^z - J_z\Delta t
238 > \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = -J_z \Delta t +
239 > \sum_{\alpha = x,y,z} K_c^\alpha  
240   \label{eq:fluxEllipsoid}
241   \end{equation}
242   The spatial extent of the {\it thermal flux ellipsoid} is governed
243 < both by a targetted value, $J_z$ as well as the instantaneous values
244 < of the kinetic energy components in the cold bin.
243 > both by the target flux, $J_z$ as well as the instantaneous values of
244 > the kinetic energy components in the cold bin.
245  
246   To satisfy an energetic flux as well as the conservation constraints,
247 < we must determine the points ${x,y,z}$ which lie on both the
248 < constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
249 < ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
250 < two ellipsoids in 3-dimensional space.
247 > we must determine the points ${x,y,z}$ that lie on both the constraint
248 > ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
249 > (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
250 > ellipsoids in 3-dimensional space.
251  
252   \begin{figure}
253   \includegraphics[width=\linewidth]{ellipsoids}
254 < \caption{Illustration from the perspective of a space having cold
255 <  slab scaling coefficients as its coordinates. Scaling points which
256 <  maintain both constant energy and constant linear momentum of the
257 <  system lie on the surface of the {\it constraint ellipsoid} while
258 <  points which generate the target momentum flux lie on the surface of
259 <  the {\it flux ellipsoid}. The velocity distributions in the cold bin
255 <  are scaled by only those points which lie on both ellipsoids.}
254 > \caption{Velocity scaling coefficients which maintain both constant
255 >  energy and constant linear momentum of the system lie on the surface
256 >  of the {\it constraint ellipsoid} while points which generate the
257 >  target momentum flux lie on the surface of the {\it flux ellipsoid}.
258 >  The velocity distributions in the cold bin are scaled by only those
259 >  points which lie on both ellipsoids.}
260   \label{ellipsoids}
261   \end{figure}
262  
263 < One may also define {\it momentum} flux (say along the $x$-direction) as:
263 > Since ellipsoids can be expressed as polynomials up to second order in
264 > each of the three coordinates, finding the the intersection points of
265 > two ellipsoids is isomorphic to finding the roots a polynomial of
266 > degree 16.  There are a number of polynomial root-finding methods in
267 > the literature, [CITATIONS NEEDED] but numerically finding the roots
268 > of high-degree polynomials is generally an ill-conditioned
269 > problem.[CITATION NEEDED] One way around this is to try to maintain
270 > velocity scalings that are {\it as isotropic as possible}.  To do
271 > this, we impose $x=y$, and to treat both the constraint and flux
272 > ellipsoids as 2-dimensional ellipses.  In reduced dimensionality, the
273 > intersecting-ellipse problem reduces to finding the roots of
274 > polynomials of degree 4.
275 >
276 > Depending on the target flux and current velocity distributions, the
277 > ellipsoids can have between 0 and 4 intersection points.  If there are
278 > no intersection points, it is not possible to satisfy the constraints
279 > while performing a non-equilibrium scaling move, and no change is made
280 > to the dynamics.  
281 >
282 > With multiple intersection points, any of the scaling points will
283 > conserve the linear momentum and kinetic energy of the system and will
284 > generate the correct target flux.  Although this method is inherently
285 > non-isotropic, the goal is still to maintain the system as close to an
286 > isotropic fluid as possible.  With this in mind, we would like the
287 > kinetic energies in the three different directions could become as
288 > close as each other as possible after each scaling.  Simultaneously,
289 > one would also like each scaling as gentle as possible, i.e. ${x,y,z
290 >  \rightarrow 1}$, in order to avoid large perturbation to the system.
291 > To do this, we pick the intersection point which maintains the scaling
292 > variables ${x=y, z}$ as well as the ratio of kinetic energies
293 > ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
294 >
295 > After the valid scaling parameters are arrived at by solving geometric
296 > intersection problems in $x, y, z$ space in order to obtain cold slab
297 > scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
298 > determine the conjugate hot slab scaling variables.
299 >
300 > \subsection{Introducing shear stress via velocity scaling}
301 > Rather than using this method to induce a thermal flux, it is possible
302 > to use the random fluctuations of the average momentum in each of the
303 > bins to induce a momentum flux.  Doing this repeatedly will create a
304 > shear stress on the system which will respond with an easily-measured
305 > strain.  The momentum flux (say along the $x$-direction) may be
306 > defined as:
307   \begin{equation}
308   (1-x) P_c^x = j_z(p_x)\Delta t
309   \label{eq:fluxPlane}
310   \end{equation}
311 < The above {\it momentum flux plane} is perpendicular to the $x$-axis,
312 < with its position governed both by a target value, $j_z(p_x)$ as well
313 < as the instantaneous value of the momentum along the $x$-direction.
311 > This {\it momentum flux plane} is perpendicular to the $x$-axis, with
312 > its position governed both by a target value, $j_z(p_x)$ as well as
313 > the instantaneous value of the momentum along the $x$-direction.
314  
315   In order to satisfy a momentum flux as well as the conservation
316   constraints, we must determine the points ${x,y,z}$ which lie on both
317   the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
318   flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
319 < ellipsoid and a plane in 3-dimensional space.  
319 > ellipsoid and a plane in 3-dimensional space.
320  
321 < In both the momentum and energy flux scenarios, valid scaling
322 < parameters are arrived at by solving geometric intersection problems
323 < in $x, y, z$ space in order to obtain cold slab scaling parameters.
324 < Once the scaling variables for the cold slab are known, the hot slab
325 < scaling has also been determined.
321 > In the case of momentum flux transfer, we also impose another
322 > constraint to set the kinetic energy transfer as zero. In another
323 > word, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$.  With
324 > one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
325 > set of quartic equations to the above kinetic energy transfer problem.
326  
327 + \section{Computational Details}
328  
329 < The following problem will be choosing an optimal set of scaling
330 < variables among the possible sets. Although this method is inherently
331 < non-isotropic, the goal is still to maintain the system as isotropic
332 < as possible. Under this consideration, one would like the kinetic
333 < energies in different directions could become as close as each other
334 < after each scaling. Simultaneously, one would also like each scaling
335 < as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
336 < large perturbation to the system. Therefore, one approach to obtain
337 < the scaling variables would be constructing an criteria function, with
290 < constraints as above equation sets, and solving the function's minimum
291 < by method like Lagrange multipliers.
329 > We have implemented this methodology in our molecular dynamics
330 > code,\cite{Meineke:2005gd} by performing the NIVS scaling moves after
331 > each MD step.  We have tested it for a variety of different
332 > situations, including homogeneous fluids (Lennard-Jones and SPC/E
333 > water), crystalline solids (EAM and Sutton-Chen models for Gold), and
334 > heterogeneous interfaces (EAM gold - SPC/E water). The last of these
335 > systems would have been very difficult to study using previous RNEMD
336 > methods, but using velocity scaling moves, we can even obtain
337 > estimates of the interfacial thermal conductivity.
338  
293 In order to save computation time, we have a different approach to a
294 relatively good set of scaling variables with much less calculation
295 than above. Here is the detail of our simplification of the problem.
296
297 In the case of kinetic energy transfer, we impose another constraint
298 ${x = y}$, into the equation sets. Consequently, there are two
299 variables left. And now one only needs to solve a set of two {\it
300  ellipses equations}. This problem would be transformed into solving
301 one quartic equation for one of the two variables. There are known
302 generic methods that solve real roots of quartic equations. Then one
303 can determine the other variable and obtain sets of scaling
304 variables. Among these sets, one can apply the above criteria to
305 choose the best set, while much faster with only a few sets to choose.
306
307 In the case of momentum flux transfer, we impose another constraint to
308 set the kinetic energy transfer as zero. In another word, we apply
309 Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
310 variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
311 of equations on the above kinetic energy transfer problem. Therefore,
312 an approach similar to the above would be sufficient for this as well.
313
314 \section{Computational Details}
339   \subsection{Lennard-Jones Fluid}
316 Our simulation consists of a series of systems. All of these
317 simulations were run with the OpenMD simulation software
318 package\cite{Meineke:2005gd} integrated with RNEMD codes.
340  
341 < A Lennard-Jones fluid system was built and tested first. In order to
342 < compare our method with swapping RNEMD, a series of simulations were
343 < performed to calculate the shear viscosity and thermal conductivity of
344 < argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
345 <  \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
346 < ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
347 < comparison between our results and others. These simulations used
348 < velocity Verlet algorithm with reduced timestep ${\tau^* =
349 <  4.6\times10^{-4}}$.
341 > 2592 Lennard-Jones atoms were placed in an orthorhombic cell
342 > ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side.  The
343 > reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
344 > direct comparison between our results and previous methods.  These
345 > simulations were carried out with a reduced timestep ${\tau^* =
346 >  4.6\times10^{-4}}$.  For the shear viscosity calculation, the mean
347 > temperature was ${T^* = k_B T/\varepsilon = 0.72}$.  Simulations were
348 > first thermalized in canonical ensemble (NVT), then equilibrated in
349 > microcanonical ensemble (NVE) before introducing any non-equilibrium
350 > method.
351  
352 < For shear viscosity calculation, the reduced temperature was ${T^* =
353 <  k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
354 < ensemble (NVT), then equilibrated in microcanonical ensemble
355 < (NVE). Establishing and stablizing momentum gradient were followed
356 < also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
357 < adopted.\cite{ISI:000080382700030} The simulation box was under
358 < periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
359 < the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
360 < most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
361 < to Tenney {\it et al.}\cite{Maginn:2010}, a series of swapping
362 < frequency were chosen. According to each result from swapping
363 < RNEMD, scaling RNEMD simulations were run with the target momentum
364 < flux set to produce a similar momentum flux, and consequently shear
365 < rate. Furthermore, various scaling frequencies can be tested for one
366 < single swapping rate. To test the temperature homogeneity in our
367 < system of swapping and scaling methods, temperatures of different
368 < dimensions in all the slabs were observed. Most of the simulations
369 < include $10^5$ steps of equilibration without imposing momentum flux,
370 < $10^5$ steps of stablization with imposing unphysical momentum
371 < transfer, and $10^6$ steps of data collection under RNEMD. For
372 < relatively high momentum flux simulations, ${5\times10^5}$ step data
373 < collection is sufficient. For some low momentum flux simulations,
374 < ${2\times10^6}$ steps were necessary.
352 > We have compared the momentum gradients established using our method
353 > to those obtained using the original M\"{u}ller-Plathe swapping
354 > algorithm.\cite{ISI:000080382700030} In both cases, the simulation box
355 > was divided into ${N = 20}$ slabs.  In the swapping algorithm, the top
356 > slab $(n = 1)$ exchanges the most negative $x$ momentum with the most
357 > positive $x$ momentum in the center slab $(n = N/2 + 1)$.  The rate at
358 > which the swapping moves are carried out defines the momentum or
359 > thermal flux between the two slabs.  In their work, Tenney {\it et
360 >  al.}\cite{Maginn:2010} found problematic behavior with large swap
361 > frequencies.
362 >
363 > According to each result from swapping RNEMD, scaling RNEMD
364 > simulations were run with the target momentum flux set to produce a
365 > similar momentum flux, and consequently shear rate.  Furthermore,
366 > various scaling frequencies can be tested for one single swapping
367 > rate. To test the temperature homogeneity in our system of swapping
368 > and scaling methods, temperatures of different dimensions in all the
369 > slabs were observed.  Most of the simulations include $10^5$ steps of
370 > equilibration without imposing momentum flux, $10^5$ steps of
371 > stablization with imposing unphysical momentum transfer, and $10^6$
372 > steps of data collection under RNEMD. For relatively high momentum
373 > flux simulations, ${5\times10^5}$ step data collection is sufficient.
374 > For some low momentum flux simulations, ${2\times10^6}$ steps were
375 > necessary.
376  
377   After each simulation, the shear viscosity was calculated in reduced
378   unit. The momentum flux was calculated with total unphysical
# Line 591 | Line 614 | shown separately in Table \ref{qscThermal} and \ref{ea
614  
615   \subsubsection{Crystal Gold}
616   Our results of gold thermal conductivity using two force fields are
617 < shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
618 < these calculations,the end and middle slabs were excluded in thermal
619 < gradient regession and only used as heat source and drain in the
620 < systems. Our yielded values using EAM force field are slightly larger
621 < than those using QSC force field. However, both series are
622 < significantly smaller than experimental value by an order of more than
623 < 100. It has been verified that this difference is mainly attributed to
624 < the lack of electron interaction representation in these force field
625 < parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
626 < force field parameters in their metal thermal conductivity
627 < calculations. The Non-Equilibrium MD method they employed in their
628 < simulations produced comparable results to ours. As Zhang {\it et
629 <  al.}\cite{ISI:000231042800044} stated, thermal conductivity values
630 < are influenced mainly by force field. Therefore, it is confident to
631 < conclude that NIVS-RNEMD is applicable to metal force field system.
632 <
633 < \begin{figure}
634 < \includegraphics[width=\linewidth]{AuGrad}
635 < \caption{Temperature gradients for thermal conductivity calculation of
636 <  crystal gold using QSC force field.}
637 < \label{AuGrad}
638 < \end{figure}
616 <
617 < \begin{table*}
618 < \begin{minipage}{\linewidth}
619 < \begin{center}
620 <
621 < \caption{Calculation results for thermal conductivity of crystal gold
622 <  using QSC force field at ${\langle T\rangle}$ = 300K at various
623 <  thermal exchange rates. Errors of calculations in parentheses. }
624 <
625 < \begin{tabular}{cc}
626 < \hline
627 < $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
628 < \hline
629 < 1.44 & 1.10(0.01)\\
630 < 2.86 & 1.08(0.02)\\
631 < 5.14 & 1.15(0.01)\\
632 < \hline
633 < \end{tabular}
634 < \label{qscThermal}
635 < \end{center}
636 < \end{minipage}
637 < \end{table*}
617 > shown in Table \ref{AuThermal}. In these calculations,the end and
618 > middle slabs were excluded in thermal gradient regession and only used
619 > as heat source and drain in the systems. Our yielded values using EAM
620 > force field are slightly larger than those using QSC force
621 > field. However, both series are significantly smaller than
622 > experimental value by a factor of more than 200. It has been verified
623 > that this difference is mainly attributed to the lack of electron
624 > interaction representation in these force field parameters. Richardson
625 > {\it et al.}\cite{Clancy:1992} used EAM force field parameters in
626 > their metal thermal conductivity calculations. The Non-Equilibrium MD
627 > method they employed in their simulations produced comparable results
628 > to ours. As Zhang {\it et al.}\cite{ISI:000231042800044} stated,
629 > thermal conductivity values are influenced mainly by force
630 > field. Another factor that affects the calculation results could be
631 > the density of the metal. After equilibration under
632 > isobaric-isothermal conditions, our crystall simulation cell expanded
633 > by the order of 1\%. Under longer lattice constant than default,
634 > lower thermal conductance would be expected. Furthermore, the result
635 > of Richardson {\it et al.} were obtained between 300K and 850K, which
636 > are significantly higher than in our simulations. Therefore, it is
637 > still confident to conclude that NIVS-RNEMD is applicable to metal
638 > force field system.
639  
639 \begin{figure}
640 \includegraphics[width=\linewidth]{eamGrad}
641 \caption{Temperature gradients for thermal conductivity calculation of
642  crystal gold using EAM force field.}
643 \label{eamGrad}
644 \end{figure}
645
640   \begin{table*}
641   \begin{minipage}{\linewidth}
642   \begin{center}
643  
644   \caption{Calculation results for thermal conductivity of crystal gold
645 <  using EAM force field at ${\langle T\rangle}$ = 300K at various
646 <  thermal exchange rates. Errors of calculations in parentheses. }
645 >  using different force fields at ${\langle T\rangle}$ = 300K at
646 >  various thermal exchange rates. Errors of calculations in parentheses.}
647  
648 < \begin{tabular}{cc}
648 > \begin{tabular}{ccc}
649   \hline
650 < $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
650 > Force Field Used & $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
651   \hline
652 < 1.24 & 1.24(0.06)\\
653 < 2.06 & 1.37(0.04)\\
654 < 2.55 & 1.41(0.03)\\
652 >    & 1.44 & 1.10(0.01)\\
653 > QSC & 2.86 & 1.08(0.02)\\
654 >    & 5.14 & 1.15(0.01)\\
655   \hline
656 +    & 1.24 & 1.24(0.06)\\
657 + EAM & 2.06 & 1.37(0.04)\\
658 +    & 2.55 & 1.41(0.03)\\
659 + \hline
660   \end{tabular}
661 < \label{eamThermal}
661 > \label{AuThermal}
662   \end{center}
663   \end{minipage}
664   \end{table*}
# Line 671 | Line 669 | surface in our system. Figure \ref{interfaceDensity} d
669   NIVS-RNEMD method were proved valid, calculation of gold/water
670   interfacial thermal conductivity was followed. It is found out that
671   the low interfacial conductance is probably due to the hydrophobic
672 < surface in our system. Figure \ref{interfaceDensity} demonstrates mass
672 > surface in our system. Figure \ref{interface} (a) demonstrates mass
673   density change along $z$-axis, which is perpendicular to the
674   gold/water interface. It is observed that water density significantly
675   decreases when approaching the surface. Under this low thermal
676   conductance, both gold and water phase have sufficient time to
677   eliminate temperature difference inside respectively (Figure
678 < \ref{interfaceGrad}). With indistinguishable temperature difference
678 > \ref{interface} b). With indistinguishable temperature difference
679   within respective phase, it is valid to assume that the temperature
680   difference between gold and water on surface would be approximately
681   the same as the difference between the gold and water phase. This
# Line 690 | Line 688 | errors than our calculation results on homogeneous sys
688   errors than our calculation results on homogeneous systems.
689  
690   \begin{figure}
691 < \includegraphics[width=\linewidth]{interfaceDensity}
692 < \caption{Density profile for interfacial thermal conductivity
693 <  simulation box. Significant water density decrease is observed on
694 <  gold surface.}
695 < \label{interfaceDensity}
691 > \includegraphics[width=\linewidth]{interface}
692 > \caption{Simulation results for Gold/Water interfacial thermal
693 >  conductivity: (a) Significant water density decrease is observed on
694 >  crystalline gold surface, which indicates low surface contact and
695 >  leads to low thermal conductance. (b) Temperature profiles for a
696 >  series of simulations. Temperatures of different slabs in the same
697 >  phase show no significant differences.}
698 > \label{interface}
699   \end{figure}
700  
700 \begin{figure}
701 \includegraphics[width=\linewidth]{interfaceGrad}
702 \caption{Temperature profiles for interfacial thermal conductivity
703  simulation box. Temperatures of different slabs in the same phase
704  show no significant difference.}
705 \label{interfaceGrad}
706 \end{figure}
707
701   \begin{table*}
702   \begin{minipage}{\linewidth}
703   \begin{center}
# Line 751 | Line 744 | Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
744  
745   \begin{tabular}{ccc}
746   \hline
747 < Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
747 > (Equilvalent) Exchange Interval (fs) & $\eta^*_{swap}$ &
748 > $\eta^*_{scale}$\\
749   \hline
750 < 20-500 & 3.64(0.05) & 3.76(0.09)\\
751 < 20-1000 & 3.52(0.16) & 3.66(0.06)\\
752 < 20-2000 & 3.72(0.05) & 3.32(0.18)\\
753 < 20-2500 & 3.42(0.06) & 3.43(0.08)\\
750 > 500  & 3.64(0.05) & 3.76(0.09)\\
751 > 1000 & 3.52(0.16) & 3.66(0.06)\\
752 > 2000 & 3.72(0.05) & 3.32(0.18)\\
753 > 2500 & 3.42(0.06) & 3.43(0.08)\\
754   \hline
755   \end{tabular}
756   \label{shearRate}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines