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 3527 by skuang, Mon Sep 21 20:54:06 2009 UTC vs.
Revision 3570 by skuang, Wed Mar 10 02:00:06 2010 UTC

# Line 56 | Line 56 | the simulation cell.\cite{MullerPlathe:1997xw,Muller-P
56   (RNEMD) obtains transport coefficients (thermal conductivity and shear
57   viscosity) in a fluid by imposing an artificial momentum flux between
58   two thin parallel slabs of material that are spatially separated in
59 < the simulation cell.\cite{MullerPlathe:1997xw,Muller-Plathe:1999ek} The
60 < artificial flux is typically created by periodically "swapping" either
59 > the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
60 > artificial flux is typically created by periodically ``swapping'' either
61   the entire momentum vector $\vec{p}$ or single components of this
62   vector ($p_x$) between molecules in each of the two slabs.  If the two
63   slabs are separated along the z coordinate, the imposed flux is either
64 < directional ($J_z(p_x)$) or isotropic ($J_z$), and the response of a
64 > directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
65   simulated system to the imposed momentum flux will typically be a
66   velocity or thermal gradient.  The transport coefficients (shear
67   viscosity and thermal conductivity) are easily obtained by assuming
68   linear response of the system,
69   \begin{eqnarray}
70 < J_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
70 > j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
71   J & = & \lambda \frac{\partial T}{\partial z}
72   \end{eqnarray}
73 < RNEMD been widely used to provide computational estimates of thermal
73 > RNEMD has been widely used to provide computational estimates of thermal
74   conductivities and shear viscosities in a wide range of materials,
75   from liquid copper to monatomic liquids to molecular fluids
76 < (e.g. ionic liquids).
76 > (e.g. ionic liquids).\cite{ISI:000246190100032}
77  
78   RNEMD is preferable in many ways to the forward NEMD methods because
79   it imposes what is typically difficult to measure (a flux or stress)
# Line 91 | Line 91 | Recently, Tenney and Maginn have discovered some probl
91   typically samples from the same manifold of states in the
92   microcanonical ensemble.
93  
94 < Recently, Tenney and Maginn have discovered some problems with the
95 < original RNEMD swap technique.  Notably, large momentum fluxes
96 < (equivalent to frequent momentum swaps between the slabs) can result
97 < in "notched", "peaked" and generally non-thermal momentum
94 > Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
95 > some problems with the original RNEMD swap technique.  Notably, large
96 > momentum fluxes (equivalent to frequent momentum swaps between the
97 > slabs) can result in ``notched'', ``peaked'' and generally non-thermal momentum
98   distributions in the two slabs, as well as non-linear thermal and
99   velocity distributions along the direction of the imposed flux ($z$).
100   Tenney and Maginn obtained reasonable limits on imposed flux and
# Line 121 | Line 121 | moves.  For molecules $\{i\}$ located within the hot s
121   hot slab.
122  
123   Rather than using momentum swaps, we use a series of velocity scaling
124 < moves.  For molecules $\{i\}$ located within the hot slab,
124 > moves.  For molecules $\{i\}$ located within the cold slab,
125   \begin{equation}
126 < \vec{v}_i \leftarrow \left( \begin{array}{c}
127 < x \\
128 < y \\
129 < z \\
126 > \vec{v}_i \leftarrow \left( \begin{array}{ccc}
127 > x & 0 & 0 \\
128 > 0 & y & 0 \\
129 > 0 & 0 & z \\
130   \end{array} \right) \cdot \vec{v}_i
131   \end{equation}
132   where ${x, y, z}$ are a set of 3 scaling variables for each of the
133   three directions in the system.  Likewise, the molecules $\{j\}$
134 < located in the cold bin will see a concomitant scaling of velocities,
134 > located in the hot slab will see a concomitant scaling of velocities,
135   \begin{equation}
136 < \vec{v}_j \leftarrow \left( \begin{array}{c}
137 < x^\prime \\
138 < y^\prime \\
139 < z^\prime \\
136 > \vec{v}_j \leftarrow \left( \begin{array}{ccc}
137 > x^\prime & 0 & 0 \\
138 > 0 & y^\prime & 0 \\
139 > 0 & 0 & z^\prime \\
140   \end{array} \right) \cdot \vec{v}_j
141   \end{equation}
142  
# Line 144 | Line 144 | P_h^\alpha + P_c^\alpha = \alpha P_h^\alpha + \alpha^\
144   ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
145   parameters together:
146   \begin{equation}
147 < P_h^\alpha + P_c^\alpha = \alpha P_h^\alpha + \alpha^\prime P_c^\alpha
147 > P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
148   \end{equation}
149   where
150 < \begin{equation}
151 < \begin{array}{rcl}
152 < P_h^\alpha & = & \sum_{i = 1}^{N_h} m_i \left[\vec{v}_i\right]_\alpha \\
153 < P_c^\alpha & = & \sum_{j = 1}^{N_c} m_j \left[\vec{v}_j\right]_\alpha
154 < \\
155 < \end{array}
150 > \begin{eqnarray}
151 > P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
152 > P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
153   \label{eq:momentumdef}
154 < \end{equation}
155 < Therefore, for each of the three directions, the cold scaling
156 < parameters are a simple function of the hot scaling parameters and
154 > \end{eqnarray}
155 > Therefore, for each of the three directions, the hot scaling
156 > parameters are a simple function of the cold scaling parameters and
157   the instantaneous linear momentum in each of the two slabs.
158   \begin{equation}
159 < \alpha^\prime = 1 + (1 - \alpha) \frac{P_h^\alpha}{P_c^\alpha}.
159 > \alpha^\prime = 1 + (1 - \alpha) p_\alpha
160   \label{eq:hotcoldscaling}
161   \end{equation}
162 + where
163 + \begin{equation}
164 + p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
165 + \end{equation}
166 + for convenience.
167  
168   Conservation of total energy also places constraints on the scaling:
169   \begin{equation}
170   \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
171 < \alpha^2 K_h^\alpha + \left(\alpha^\prime\right)^2 K_c^\alpha.
171 > \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
172   \end{equation}
173   where the kinetic energies, $K_h^\alpha$ and $K_c^\alpha$, are computed
174   for each of the three directions in a similar manner to the linear momenta
175   (Eq. \ref{eq:momentumdef}).  Substituting in the expressions for the
176 < cold scaling parameters ($\alpha^\prime$) from
177 < Eq. (\ref{eq:hotcoldscaling}), we obtain the {\it constraint ellipsoid}:
176 > hot scaling parameters ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}),
177 > we obtain the {\it constraint ellipsoid equation}:
178   \begin{equation}
179 < C_x (1+x)^2 + C_y (1+y)^2 + C_z (1+z)^2 = 0,
179 > \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
180   \label{eq:constraintEllipsoid}
181   \end{equation}
182   where the constants are obtained from the instantaneous values of the
183   linear momenta and kinetic energies for the hot and cold slabs,
184 < \begin{equation}
185 < C_\alpha = \left( K_h^\alpha + K_c^\alpha
186 <  \left(\frac{P_h^\alpha}{P_c^\alpha}\right)^2\right)
184 > \begin{eqnarray}
185 > a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
186 >  \left(p_\alpha\right)^2\right) \\
187 > b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
188 > c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
189   \label{eq:constraintEllipsoidConsts}
190 < \end{equation}
191 < This ellipsoid defines the set of hot slab scaling parameters which can be
192 < applied while preserving both linear momentum in all three directions
193 < as well as total energy.
190 > \end{eqnarray}
191 > This ellipsoid equation defines the set of cold slab scaling
192 > parameters which can be applied while preserving both linear momentum
193 > in all three directions as well as kinetic energy.
194  
195   The goal of using velocity scaling variables is to transfer linear
196   momentum or kinetic energy from the cold slab to the hot slab.  If the
197   hot and cold slabs are separated along the z-axis, the energy flux is
198 < given simply by the increase in kinetic energy of the hot bin:
198 > given simply by the decrease in kinetic energy of the cold bin:
199   \begin{equation}
200 < (x^2-1) K_h^x  + (y^2-1) K_h^y + (z^2-1) K_h^z = J_z
200 > (1-x^2) K_c^x  + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
201   \end{equation}
202   The expression for the energy flux can be re-written as another
203   ellipsoid centered on $(x,y,z) = 0$:
204   \begin{equation}
205 < x^2 K_h^x + y^2 K_h^y + z^2 K_h^z = (J_z + K_h^x + K_h^y + K_h^z)
205 > 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
206   \label{eq:fluxEllipsoid}
207   \end{equation}
208 < The spatial extent of the {\it flux ellipsoid} is governed both by a
209 < targetted value, $J_z$ as well as the instantaneous values of the
210 < kinetic energy components in the hot bin.
208 > The spatial extent of the {\it flux ellipsoid equation} is governed
209 > both by a targetted value, $J_z$ as well as the instantaneous values of the
210 > kinetic energy components in the cold bin.
211  
212   To satisfy an energetic flux as well as the conservation constraints,
213   it is sufficient to determine the points ${x,y,z}$ which lie on both
214   the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
215   flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
216 < the two ellipsoids in 3-space.
216 > the two ellipsoids in 3-dimensional space.
217  
218 + \begin{figure}
219 + \includegraphics[width=\linewidth]{ellipsoids}
220 + \caption{Scaling points which maintain both constant energy and
221 +  constant linear momentum of the system lie on the surface of the
222 +  {\it constraint ellipsoid} while points which generate the target
223 +  momentum flux lie on the surface of the {\it flux ellipsoid}.  The
224 +  velocity distributions in the hot bin are scaled by only those
225 +  points which lie on both ellipsoids.}
226 + \label{ellipsoids}
227 + \end{figure}
228  
229   One may also define momentum flux (say along the x-direction) as:
230   \begin{equation}
231 < (x-1) P_h^x  = j_z(p_x)
231 > (1-x) P_c^x = j_z(p_x)\Delta t
232 > \label{eq:fluxPlane}
233 > \end{equation}
234 > The above {\it flux equation} is essentially a plane which is
235 > perpendicular to the x-axis, with its position governed both by a
236 > targetted value, $j_z(p_x)$ as well as the instantaneous value of the
237 > momentum along the x-direction.
238 >
239 > Similarly, to satisfy a momentum flux as well as the conservation
240 > constraints, it is sufficient to determine the points ${x,y,z}$ which
241 > lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
242 > and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
243 > an ellipsoid and a plane in 3-dimensional space.
244 >
245 > To summarize, by solving respective equation sets, one can determine
246 > possible sets of scaling variables for cold slab. And corresponding
247 > sets of scaling variables for hot slab can be determine as well.
248 >
249 > The following problem will be choosing an optimal set of scaling
250 > variables among the possible sets. Although this method is inherently
251 > non-isotropic, the goal is still to maintain the system as isotropic
252 > as possible. Under this consideration, one would like the kinetic
253 > energies in different directions could become as close as each other
254 > after each scaling. Simultaneously, one would also like each scaling
255 > as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
256 > large perturbation to the system. Therefore, one approach to obtain the
257 > scaling variables would be constructing an criteria function, with
258 > constraints as above equation sets, and solving the function's minimum
259 > by method like Lagrange multipliers.
260 >
261 > In order to save computation time, we have a different approach to a
262 > relatively good set of scaling variables with much less calculation
263 > than above. Here is the detail of our simplification of the problem.
264 >
265 > In the case of kinetic energy transfer, we impose another constraint
266 > ${x = y}$, into the equation sets. Consequently, there are two
267 > variables left. And now one only needs to solve a set of two {\it
268 >  ellipses equations}. This problem would be transformed into solving
269 > one quartic equation for one of the two variables. There are known
270 > generic methods that solve real roots of quartic equations. Then one
271 > can determine the other variable and obtain sets of scaling
272 > variables. Among these sets, one can apply the above criteria to
273 > choose the best set, while much faster with only a few sets to choose.
274 >
275 > In the case of momentum flux transfer, we impose another constraint to
276 > set the kinetic energy transfer as zero. In another word, we apply
277 > Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
278 > variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
279 > of equations on the above kinetic energy transfer problem. Therefore,
280 > an approach similar to the above would be sufficient for this as well.
281 >
282 > \section{Computational Details}
283 > Our simulation consists of a series of systems. All of these
284 > simulations were run with the OpenMD simulation software
285 > package\cite{Meineke:2005gd} integrated with RNEMD methods.
286 >
287 > A Lennard-Jones fluid system was built and tested first. In order to
288 > compare our method with swapping RNEMD, a series of simulations were
289 > performed to calculate the shear viscosity and thermal conductivity of
290 > argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
291 >  \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
292 > ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
293 > comparison between our results and others. These simulations used
294 > velocity Verlet algorithm with reduced timestep ${\tau^* =
295 >  4.6\times10^{-4}}$.
296 >
297 > For shear viscosity calculation, the reduced temperature was ${T^* =
298 >  k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
299 > ensemble (NVT), then equilibrated in microcanonical ensemble
300 > (NVE). Establishing and stablizing momentum gradient were followed
301 > also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
302 > adopted.\cite{ISI:000080382700030} The simulation box was under
303 > periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
304 > the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
305 > most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
306 > to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
307 > frequency were chosen. According to each result from swapping
308 > RNEMD, scaling RNEMD simulations were run with the target momentum
309 > flux set to produce a similar momentum flux and shear
310 > rate. Furthermore, various scaling frequencies can be tested for one
311 > single swapping rate. To compare the performance between swapping and
312 > scaling method, temperatures of different dimensions in all the slabs
313 > were observed. Most of the simulations include $10^5$ steps of
314 > equilibration without imposing momentum flux, $10^5$ steps of
315 > stablization with imposing momentum transfer, and $10^6$ steps of data
316 > collection under RNEMD. For relatively high momentum flux simulations,
317 > ${5\times10^5}$ step data collection is sufficient. For some low momentum
318 > flux simulations, ${2\times10^6}$ steps were necessary.
319 >
320 > After each simulation, the shear viscosity was calculated in reduced
321 > unit. The momentum flux was calculated with total unphysical
322 > transferred momentum ${P_x}$ and data collection time $t$:
323 > \begin{equation}
324 > j_z(p_x) = \frac{P_x}{2 t L_x L_y}
325 > \end{equation}
326 > And the velocity gradient ${\langle \partial v_x /\partial z \rangle}$
327 > can be obtained by a linear regression of the velocity profile. From
328 > the shear viscosity $\eta$ calculated with the above parameters, one
329 > can further convert it into reduced unit ${\eta^* = \eta \sigma^2
330 >  (\varepsilon  m)^{-1/2}}$.
331 >
332 > For thermal conductivity calculation, simulations were first run under
333 > reduced temperature ${T^* = 0.72}$ in NVE ensemble. Muller-Plathe's
334 > algorithm was adopted in the swapping method. Under identical
335 > simulation box parameters, in each swap, the top slab exchange the
336 > molecule with least kinetic energy with the molecule in the center
337 > slab with most kinetic energy, unless this ``coldest'' molecule in the
338 > ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the ``cold''
339 > slab. According to swapping RNEMD results, target energy flux for
340 > scaling RNEMD simulations can be set. Also, various scaling
341 > frequencies can be tested for one target energy flux. To compare the
342 > performance between swapping and scaling method, distributions of
343 > velocity and speed in different slabs were observed.
344 >
345 > For each swapping rate, thermal conductivity was calculated in reduced
346 > unit. The energy flux was calculated similarly to the momentum flux,
347 > with total unphysical transferred energy ${E_{total}}$ and data collection
348 > time $t$:
349 > \begin{equation}
350 > J_z = \frac{E_{total}}{2 t L_x L_y}
351   \end{equation}
352 + And the temperature gradient ${\langle\partial T/\partial z\rangle}$
353 + can be obtained by a linear regression of the temperature
354 + profile. From the thermal conductivity $\lambda$ calculated, one can
355 + further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
356 +  m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
357  
358 + Another series of our simulation is to calculate the interfacial
359 + thermal conductivity of a Au/H${_2}$O system. Respective calculations of
360 + water (SPC/E) and gold (QSC) thermal conductivity were performed and
361 + compared with current results to ensure the validity of
362 + NIVS-RNEMD. After that, the mixture system was simulated.
363  
364 + \section{Results And Discussion}
365 + \subsection{Shear Viscosity}
366 + Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
367 + produced comparable shear viscosity to swap RNEMD method. In Table
368 + \ref{shearRate}, the names of the calculated samples are devided into
369 + two parts. The first number refers to total slabs in one simulation
370 + box. The second number refers to the swapping interval in swap method, or
371 + in scale method the equilvalent swapping interval that the same
372 + momentum flux would theoretically result in swap method. All the scale
373 + method results were from simulations that had a scaling interval of 10
374 + time steps. The average molecular momentum gradients of these samples
375 + are shown in Figure \ref{shearGrad}.
376  
377 + \begin{table*}
378 + \begin{minipage}{\linewidth}
379 + \begin{center}
380  
381 + \caption{Calculation results for shear viscosity of Lennard-Jones
382 +  fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
383 +  methods at various momentum exchange rates. Results in reduced
384 +  unit. Errors of calculations in parentheses. }
385  
386 + \begin{tabular}{ccc}
387 + \hline
388 + Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
389 + \hline
390 + 20-500 & 3.64(0.05) & 3.76(0.09)\\
391 + 20-1000 & 3.52(0.16) & 3.66(0.06)\\
392 + 20-2000 & 3.72(0.05) & 3.32(0.18)\\
393 + 20-2500 & 3.42(0.06) & 3.43(0.08)\\
394 + \hline
395 + \end{tabular}
396 + \label{shearRate}
397 + \end{center}
398 + \end{minipage}
399 + \end{table*}
400 +
401 + \begin{figure}
402 + \includegraphics[width=\linewidth]{shearGrad}
403 + \caption{Average momentum gradients of shear viscosity simulations}
404 + \label{shearGrad}
405 + \end{figure}
406 +
407 + \begin{figure}
408 + \includegraphics[width=\linewidth]{shearTempScale}
409 + \caption{Temperature profile for scaling RNEMD simulation.}
410 + \label{shearTempScale}
411 + \end{figure}
412 + However, observations of temperatures along three dimensions show that
413 + inhomogeneity occurs in scaling RNEMD simulations, particularly in the
414 + two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
415 + relatively large imposed momentum flux, the temperature difference among $x$
416 + and the other two dimensions was significant. This would result from the
417 + algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
418 + momentum gradient is set up, $P_c^x$ would be roughly stable
419 + ($<0$). Consequently, scaling factor $x$ would most probably larger
420 + than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
421 + keep increase after most scaling steps. And if there is not enough time
422 + for the kinetic energy to exchange among different dimensions and
423 + different slabs, the system would finally build up temperature
424 + (kinetic energy) difference among the three dimensions. Also, between
425 + $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
426 + are closer to neighbor slabs. This is due to momentum transfer along
427 + $z$ dimension between slabs.
428 +
429 + Although results between scaling and swapping methods are comparable,
430 + the inherent temperature inhomogeneity even in relatively low imposed
431 + exchange momentum flux simulations makes scaling RNEMD method less
432 + attractive than swapping RNEMD in shear viscosity calculation.
433 +
434 + \subsection{Thermal Conductivity}
435 +
436 + Our thermal conductivity calculations also show that scaling method results
437 + agree with swapping method. Table \ref{thermal} lists our simulation
438 + results with similar manner we used in shear viscosity
439 + calculation. All the data reported from scaling method were obtained
440 + by simulations of 10-step exchange frequency, and the target exchange
441 + kinetic energy were set to produce equivalent kinetic energy flux as
442 + in swapping method. Figure \ref{thermalGrad} exhibits similar thermal
443 + gradients of respective similar kinetic energy flux.
444 +
445 + \begin{table*}
446 + \begin{minipage}{\linewidth}
447 + \begin{center}
448 +
449 + \caption{Calculation results for thermal conductivity of Lennard-Jones
450 +  fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
451 +  swap and scale methods at various kinetic energy exchange rates. Results
452 +  in reduced unit. Errors of calculations in parentheses.}
453 +
454 + \begin{tabular}{ccc}
455 + \hline
456 + Series & $\lambda^*_{swap}$ & $\lambda^*_{scale}$\\
457 + \hline
458 + 20-250  & 7.03(0.34) & 7.30(0.10)\\
459 + 20-500  & 7.03(0.14) & 6.95(0.09)\\
460 + 20-1000 & 6.91(0.42) & 7.19(0.07)\\
461 + 20-2000 & 7.52(0.15) & 7.19(0.28)\\
462 + \hline
463 + \end{tabular}
464 + \label{thermal}
465 + \end{center}
466 + \end{minipage}
467 + \end{table*}
468 +
469 + \begin{figure}
470 + \includegraphics[width=\linewidth]{thermalGrad}
471 + \caption{Temperature gradients of thermal conductivity simulations}
472 + \label{thermalGrad}
473 + \end{figure}
474 +
475 + During these simulations, molecule velocities were recorded in 1000 of
476 + all the snapshots. These velocity data were used to produce histograms
477 + of velocity and speed distribution in different slabs. From these
478 + histograms, it is observed that with increasing unphysical kinetic
479 + energy flux, speed and velocity distribution of molecules in slabs
480 + where swapping occured could deviate from Maxwell-Boltzmann
481 + distribution. Figure \ref{histSwap} indicates how these distributions
482 + deviate from ideal condition. In high temperature slabs, probability
483 + density in low speed is confidently smaller than ideal distribution;
484 + in low temperature slabs, probability density in high speed is smaller
485 + than ideal. This phenomenon is observable even in our relatively low
486 + swapping rate simulations. And this deviation could also leads to
487 + deviation of distribution of velocity in various dimensions. One
488 + feature of these deviated distribution is that in high temperature
489 + slab, the ideal Gaussian peak was changed into a relatively flat
490 + plateau; while in low temperature slab, that peak appears sharper.
491 +
492 + \begin{figure}
493 + \includegraphics[width=\linewidth]{histSwap}
494 + \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
495 + \label{histSwap}
496 + \end{figure}
497 +
498 + \begin{figure}
499 + \includegraphics[width=\linewidth]{histScale}
500 + \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
501 + \label{histScale}
502 + \end{figure}
503 +
504 + \subsection{Interfaciel Thermal Conductivity}
505 +
506 + \begin{figure}
507 + \includegraphics[width=\linewidth]{spceGrad}
508 + \caption{Temperature gradients for SPC/E water thermal conductivity.}
509 + \label{spceGrad}
510 + \end{figure}
511 +
512 + \begin{table*}
513 + \begin{minipage}{\linewidth}
514 + \begin{center}
515 +
516 + \caption{Calculation results for thermal conductivity of SPC/E water
517 +  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
518 +  calculations in parentheses. }
519 +
520 + \begin{tabular}{cccc}
521 + \hline
522 + $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
523 + & This work & Previous simulations$^a$ & Experiment$^b$\\
524 + \hline
525 + 0.3 & 0.82()  & 0.784 & 0.64\\
526 + 0.8 & 0.770() & 0.730\\
527 + 1.5 & 0.813() & \\
528 + \hline
529 + \end{tabular}
530 + \label{spceThermal}
531 + \end{center}
532 + \end{minipage}
533 + \end{table*}
534 +
535 +
536 + \begin{figure}
537 + \includegraphics[width=\linewidth]{AuGrad}
538 + \caption{Temperature gradients for crystal gold thermal conductivity.}
539 + \label{AuGrad}
540 + \end{figure}
541 +
542 + \begin{table*}
543 + \begin{minipage}{\linewidth}
544 + \begin{center}
545 +
546 + \caption{Calculation results for thermal conductivity of crystal gold
547 +  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
548 +  calculations in parentheses. }
549 +
550 + \begin{tabular}{ccc}
551 + \hline
552 + $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
553 + & This work & Previous simulations$^a$ \\
554 + \hline
555 + 1.4 & 1.10() & \\
556 + 2.8 & 1.08() & \\
557 + 5.1 & 1.15() & \\
558 + \hline
559 + \end{tabular}
560 + \label{AuThermal}
561 + \end{center}
562 + \end{minipage}
563 + \end{table*}
564 +
565 +
566   \section{Acknowledgments}
567   Support for this project was provided by the National Science
568   Foundation under grant CHE-0848243. Computational time was provided by

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines