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 3577 by skuang, Thu Mar 18 00:53:20 2010 UTC vs.
Revision 3587 by skuang, Wed Apr 14 19:56:44 2010 UTC

# Line 38 | Line 38 | Notre Dame, Indiana 46556}
38   \begin{doublespace}
39  
40   \begin{abstract}
41 <
41 >  We present a new method for introducing stable non-equilibrium
42 >  velocity and temperature distributions in molecular dynamics
43 >  simulations of heterogeneous systems.  This method extends some
44 >  earlier Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods
45 >  which use momentum exchange swapping moves that can create
46 >  non-thermal velocity distributions (and which are difficult to use
47 >  for interfacial calculations).  By using non-isotropic velocity
48 >  scaling (NIVS) on the molecules in specific regions of a system, it
49 >  is possible to impose momentum or thermal flux between regions of a
50 >  simulation and stable thermal and momentum gradients can then be
51 >  established.  The scaling method we have developed conserves the
52 >  total linear momentum and total energy of the system.  To test the
53 >  methods, we have computed the thermal conductivity of model liquid
54 >  and solid systems as well as the interfacial thermal conductivity of
55 >  a metal-water interface.  We find that the NIVS-RNEMD improves the
56 >  problematic velocity distributions that develop in other RNEMD
57 >  methods.
58   \end{abstract}
59  
60   \newpage
# Line 49 | Line 65 | Notre Dame, Indiana 46556}
65   %                          BODY OF TEXT
66   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67  
52
53
68   \section{Introduction}
69   The original formulation of Reverse Non-equilibrium Molecular Dynamics
70   (RNEMD) obtains transport coefficients (thermal conductivity and shear
71   viscosity) in a fluid by imposing an artificial momentum flux between
72   two thin parallel slabs of material that are spatially separated in
73   the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 < artificial flux is typically created by periodically ``swapping'' either
75 < the entire momentum vector $\vec{p}$ or single components of this
76 < vector ($p_x$) between molecules in each of the two slabs.  If the two
77 < slabs are separated along the $z$ coordinate, the imposed flux is either
78 < directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
79 < simulated system to the imposed momentum flux will typically be a
80 < velocity or thermal gradient (Fig. \ref{thermalDemo}).  The transport
81 < coefficients (shear viscosity and thermal conductivity) are easily
82 < obtained by assuming linear response of the system,
74 > artificial flux is typically created by periodically ``swapping''
75 > either the entire momentum vector $\vec{p}$ or single components of
76 > this vector ($p_x$) between molecules in each of the two slabs.  If
77 > the two slabs are separated along the $z$ coordinate, the imposed flux
78 > is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
79 > response of a simulated system to the imposed momentum flux will
80 > typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
81 > The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
82 > easily obtained by assuming linear response of the system,
83   \begin{eqnarray}
84   j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85   J_z & = & \lambda \frac{\partial T}{\partial z}
# Line 73 | Line 87 | from liquid copper to monatomic liquids to molecular f
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{ISI:000246190100032}
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}
91  
92   \begin{figure}
93   \includegraphics[width=\linewidth]{thermalDemo}
94 < \caption{Demostration of thermal gradient estalished by RNEMD
95 <  method. Physical thermal flow directs from high temperature region
96 <  to low temperature region. Unphysical thermal transfer counteracts
97 <  it and maintains a steady thermal gradient.}
94 > \caption{RNEMD methods impose an unphysical transfer of momentum or
95 >  kinetic energy between a ``hot'' slab and a ``cold'' slab in the
96 >  simulation box.  The molecular system responds to this imposed flux
97 >  by generating a momentum or temperature gradient.  The slope of the
98 >  gradient can then be used to compute transport properties (e.g.
99 >  shear viscosity and thermal conductivity).}
100   \label{thermalDemo}
101   \end{figure}
102  
103 < RNEMD is preferable in many ways to the forward NEMD methods because
104 < it imposes what is typically difficult to measure (a flux or stress)
105 < and it is typically much easier to compute momentum gradients or
106 < strains (the response).  For similar reasons, RNEMD is also preferable
107 < to slowly-converging equilibrium methods for measuring thermal
108 < conductivity and shear viscosity (using Green-Kubo relations or the
109 < Helfand moment approach of Viscardy {\it et
103 > RNEMD is preferable in many ways to the forward NEMD methods
104 > [CITATIONS NEEDED] because it imposes what is typically difficult to measure
105 > (a flux or stress) and it is typically much easier to compute momentum
106 > gradients or strains (the response).  For similar reasons, RNEMD is
107 > also preferable to slowly-converging equilibrium methods for measuring
108 > thermal conductivity and shear viscosity (using Green-Kubo relations
109 > [CITATIONS NEEDED] or the Helfand moment approach of Viscardy {\it et
110    al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
111   computing difficult to measure quantities.
112  
# Line 114 | Line 130 | develop the method for determining the scaling constra
130   (conservation of linear momentum and total energy, compatibility with
131   periodic boundary conditions) while establishing true thermal
132   distributions in each of the two slabs.  In the next section, we
133 < develop the method for determining the scaling constraints.  We then
133 > present the method for determining the scaling constraints.  We then
134   test the method on both single component, multi-component, and
135   non-isotropic mixtures and show that it is capable of providing
136   reasonable estimates of the thermal conductivity and shear viscosity
137   in these cases.
138  
139   \section{Methodology}
140 < We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
141 < system is partitioned into a series of thin slabs along a particular
140 > We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
141 > periodic system is partitioned into a series of thin slabs along one
142   axis ($z$).  One of the slabs at the end of the periodic box is
143   designated the ``hot'' slab, while the slab in the center of the box
144   is designated the ``cold'' slab.  The artificial momentum flux will be
# Line 130 | Line 146 | moves.  For molecules $\{i\}$ located within the cold
146   hot slab.
147  
148   Rather than using momentum swaps, we use a series of velocity scaling
149 < moves.  For molecules $\{i\}$ located within the cold slab,
149 > moves.  For molecules $\{i\}$  located within the cold slab,
150   \begin{equation}
151   \vec{v}_i \leftarrow \left( \begin{array}{ccc}
152   x & 0 & 0 \\
# Line 150 | Line 166 | Conservation of linear momentum in each of the three d
166   \end{equation}
167  
168   Conservation of linear momentum in each of the three directions
169 < ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
169 > ($\alpha = x,y,z$) ties the values of the hot and cold scaling
170   parameters together:
171   \begin{equation}
172   P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
# Line 184 | Line 200 | Substituting in the expressions for the hot scaling pa
200   similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
201   Substituting in the expressions for the hot scaling parameters
202   ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
203 < {\it constraint ellipsoid equation}:
203 > {\it constraint ellipsoid}:
204   \begin{equation}
205   \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
206   \label{eq:constraintEllipsoid}
# Line 198 | Line 214 | This ellipsoid equation defines the set of cold slab s
214   c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
215   \label{eq:constraintEllipsoidConsts}
216   \end{eqnarray}
217 < This ellipsoid equation defines the set of cold slab scaling
218 < parameters which can be applied while preserving both linear momentum
219 < in all three directions as well as kinetic energy.
217 > This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
218 > cold slab scaling parameters which can be applied while preserving
219 > both linear momentum in all three directions as well as total kinetic
220 > energy.
221  
222   The goal of using velocity scaling variables is to transfer linear
223   momentum or kinetic energy from the cold slab to the hot slab.  If the
# Line 215 | Line 232 | The spatial extent of the {\it thermal flux ellipsoid
232   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
233   \label{eq:fluxEllipsoid}
234   \end{equation}
235 < The spatial extent of the {\it thermal flux ellipsoid equation} is
236 < governed both by a targetted value, $J_z$ as well as the instantaneous
237 < values of the kinetic energy components in the cold bin.
235 > The spatial extent of the {\it thermal flux ellipsoid} is governed
236 > both by a targetted value, $J_z$ as well as the instantaneous values
237 > of the kinetic energy components in the cold bin.
238  
239   To satisfy an energetic flux as well as the conservation constraints,
240 < it is sufficient to determine the points ${x,y,z}$ which lie on both
241 < the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
242 < flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
243 < the two ellipsoids in 3-dimensional space.
240 > we must determine the points ${x,y,z}$ which lie on both the
241 > constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
242 > ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
243 > two ellipsoids in 3-dimensional space.
244  
245   \begin{figure}
246   \includegraphics[width=\linewidth]{ellipsoids}
# Line 236 | Line 253 | One may also define momentum flux (say along the $x$-d
253   \label{ellipsoids}
254   \end{figure}
255  
256 < One may also define momentum flux (say along the $x$-direction) as:
256 > One may also define {\it momentum} flux (say along the $x$-direction) as:
257   \begin{equation}
258   (1-x) P_c^x = j_z(p_x)\Delta t
259   \label{eq:fluxPlane}
260   \end{equation}
261 < The above {\it momentum flux equation} is essentially a plane which is
262 < perpendicular to the $x$-axis, with its position governed both by a
263 < target value, $j_z(p_x)$ as well as the instantaneous value of the
247 < momentum along the $x$-direction.
261 > The above {\it momentum flux plane} is perpendicular to the $x$-axis,
262 > with its position governed both by a target value, $j_z(p_x)$ as well
263 > as the instantaneous value of the momentum along the $x$-direction.
264  
265 < Similarly, to satisfy a momentum flux as well as the conservation
266 < constraints, it is sufficient to determine the points ${x,y,z}$ which
267 < lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
268 < and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
269 < an ellipsoid and a plane in 3-dimensional space.
265 > In order to satisfy a momentum flux as well as the conservation
266 > constraints, we must determine the points ${x,y,z}$ which lie on both
267 > the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
268 > flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
269 > ellipsoid and a plane in 3-dimensional space.  
270  
271 < To summarize, by solving respective equation sets, one can determine
272 < possible sets of scaling variables for cold slab. And corresponding
273 < sets of scaling variables for hot slab can be determine as well.
271 > In both the momentum and energy flux scenarios, valid scaling
272 > parameters are arrived at by solving geometric intersection problems
273 > in $x, y, z$ space in order to obtain cold slab scaling parameters.
274 > Once the scaling variables for the cold slab are known, the hot slab
275 > scaling has also been determined.
276  
277 +
278   The following problem will be choosing an optimal set of scaling
279   variables among the possible sets. Although this method is inherently
280   non-isotropic, the goal is still to maintain the system as isotropic
# Line 263 | Line 282 | large perturbation to the system. Therefore, one appro
282   energies in different directions could become as close as each other
283   after each scaling. Simultaneously, one would also like each scaling
284   as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
285 < large perturbation to the system. Therefore, one approach to obtain the
286 < scaling variables would be constructing an criteria function, with
285 > large perturbation to the system. Therefore, one approach to obtain
286 > the scaling variables would be constructing an criteria function, with
287   constraints as above equation sets, and solving the function's minimum
288   by method like Lagrange multipliers.
289  
# Line 374 | Line 393 | liquid water (SPC/E) and crystal gold (QSC) thermal co
393   \subsection{ Water / Metal Thermal Conductivity}
394   Another series of our simulation is the calculation of interfacial
395   thermal conductivity of a Au/H$_2$O system. Respective calculations of
396 < liquid water (SPC/E) and crystal gold (QSC) thermal conductivity were
397 < performed and compared with current results to ensure the validity of
398 < NIVS-RNEMD. After that, a mixture system was simulated.
396 > liquid water (Extended Simple Point Charge model) and crystal gold
397 > thermal conductivity were performed and compared with current results
398 > to ensure the validity of NIVS-RNEMD. After that, a mixture system was
399 > simulated.
400  
401   For thermal conductivity calculation of bulk water, a simulation box
402   consisting of 1000 molecules were first equilibrated under ambient
# Line 388 | Line 408 | protocol. The face centered cubic crystal simulation b
408   process was similar to Lennard-Jones fluid system.
409  
410   Thermal conductivity calculation of bulk crystal gold used a similar
411 < protocol. The face centered cubic crystal simulation box consists of
411 > protocol. Two types of force field parameters, Embedded Atom Method
412 > (EAM) and Quantum Sutten-Chen (QSC) force field were used
413 > respectively. The face-centered cubic crystal simulation box consists of
414   2880 Au atoms. The lattice was first allowed volume change to relax
415   under ambient temperature and pressure. Equilibrations in canonical and
416   microcanonical ensemble were followed in order. With the simulation
# Line 463 | Line 485 | all the snapshots. These velocity data were used to pr
485   \end{figure}
486  
487   During these simulations, molecule velocities were recorded in 1000 of
488 < all the snapshots. These velocity data were used to produce histograms
489 < of velocity and speed distribution in different slabs. From these
490 < histograms, it is observed that with increasing unphysical kinetic
491 < energy flux, speed and velocity distribution of molecules in slabs
492 < where swapping occured could deviate from Maxwell-Boltzmann
493 < distribution. Figure \ref{histSwap} indicates how these distributions
494 < deviate from ideal condition. In high temperature slabs, probability
495 < density in low speed is confidently smaller than ideal distribution;
496 < in low temperature slabs, probability density in high speed is smaller
497 < than ideal. This phenomenon is observable even in our relatively low
498 < swapping rate simulations. And this deviation could also leads to
499 < deviation of distribution of velocity in various dimensions. One
500 < feature of these deviated distribution is that in high temperature
501 < slab, the ideal Gaussian peak was changed into a relatively flat
502 < plateau; while in low temperature slab, that peak appears sharper.
488 > all the snapshots of one single data collection process. These
489 > velocity data were used to produce histograms of velocity and speed
490 > distribution in different slabs. From these histograms, it is observed
491 > that under relatively high unphysical kinetic energy flux, speed and
492 > velocity distribution of molecules in slabs where swapping occured
493 > could deviate from Maxwell-Boltzmann distribution. Figure
494 > \ref{histSwap} illustrates how these distributions deviate from an
495 > ideal distribution. In high temperature slab, probability density in
496 > low speed is confidently smaller than ideal curve fit; in low
497 > temperature slab, probability density in high speed is smaller than
498 > ideal, while larger than ideal in low speed. This phenomenon is more
499 > obvious in our high swapping rate simulations. And this deviation
500 > could also leads to deviation of distribution of velocity in various
501 > dimensions. One feature of these deviated distribution is that in high
502 > temperature slab, the ideal Gaussian peak was changed into a
503 > relatively flat plateau; while in low temperature slab, that peak
504 > appears sharper. This problem is rooted in the mechanism of the
505 > swapping method. Continually depleting low (high) speed particles in
506 > the high (low) temperature slab could not be complemented by
507 > diffusions of low (high) speed particles from neighbor slabs, unless
508 > in suffciently low swapping rate. Simutaneously, surplus low speed
509 > particles in the low temperature slab do not have sufficient time to
510 > diffuse to neighbor slabs. However, thermal exchange rate should reach
511 > a minimum level to produce an observable thermal gradient under noise
512 > interference. Consequently, swapping RNEMD has a relatively narrow
513 > choice of swapping rate to satisfy these above restrictions.
514  
515   \begin{figure}
516   \includegraphics[width=\linewidth]{histSwap}
517 < \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
517 > \caption{Speed distribution for thermal conductivity using swapping
518 >  RNEMD. Shown is from the simulation with 250 fs exchange interval.}
519   \label{histSwap}
520   \end{figure}
521  
522 + Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
523 + curve fit (Figure \ref{histScale}). Essentially, after scaling, a
524 + Gaussian distribution function would remain Gaussian. Although a
525 + single scaling is non-isotropic in all three dimensions, our scaling
526 + coefficient criteria could help maintian the scaling region as
527 + isotropic as possible. On the other hand, scaling coefficients are
528 + preferred to be as close to 1 as possible, which also helps minimize
529 + the difference among different dimensions. This is possible if scaling
530 + interval and one-time thermal transfer energy are well
531 + chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
532 + thermal flux as the previous RNEMD method without large perturbation
533 + to the distribution of velocity and speed in the exchange regions.
534 +
535   \begin{figure}
536   \includegraphics[width=\linewidth]{histScale}
537 < \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
537 > \caption{Speed distribution for thermal conductivity using scaling
538 >  RNEMD. Shown is from the simulation with an equilvalent thermal flux
539 >  as an 250 fs exchange interval swapping simulation.}
540   \label{histScale}
541   \end{figure}
542  
543   \subsubsection{SPC/E Water}
544   Our results of SPC/E water thermal conductivity are comparable to
545   Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
546 < previous swapping RNEMD method for their calculation. Our simulations
547 < were able to produce a similar temperature gradient to their
548 < system. However, the average temperature of our system is 300K, while
549 < theirs is 318K, which would be attributed for part of the difference
550 < between the two series of results. Both methods yields values in
551 < agreement with experiment. And this shows the applicability of our
552 < method to multi-atom molecular system.
546 > previous swapping RNEMD method for their calculation. Bedrov {\it et
547 >  al.}\cite{ISI:000090151400044} argued that exchange of the molecule
548 > center-of-mass velocities instead of single atom velocities in a
549 > molecule conserves the total kinetic energy and linear momentum. This
550 > principle is adopted in our simulations. Scaling is applied to the
551 > velocities of the rigid bodies of SPC/E model water molecules, instead
552 > of each hydrogen and oxygen atoms in relevant water molecules. As
553 > shown in Figure \ref{spceGrad}, temperature gradients were established
554 > similar to their system. However, the average temperature of our
555 > system is 300K, while theirs is 318K, which would be attributed for
556 > part of the difference between the final calculation results (Table
557 > \ref{spceThermal}). Both methods yields values in agreement with
558 > experiment. And this shows the applicability of our method to
559 > multi-atom molecular system.
560  
561   \begin{figure}
562   \includegraphics[width=\linewidth]{spceGrad}
# Line 533 | Line 589 | Our results of gold thermal conductivity used QSC forc
589   \end{table*}
590  
591   \subsubsection{Crystal Gold}
592 < Our results of gold thermal conductivity used QSC force field are
593 < shown in Table \ref{AuThermal}. Although our calculation is smaller
594 < than experimental value by an order of more than 100, this difference
595 < is mainly attributed to the lack of electron interaction
596 < representation in our force field parameters. Richardson {\it et
597 <  al.}\cite{ISI:A1992HX37800010} used similar force field parameters
598 < in their metal thermal conductivity calculations. The EMD method they
599 < employed in their simulations produced comparable results to
600 < ours. Therefore, it is confident to conclude that NIVS-RNEMD is
601 < applicable to metal force field system.
592 > Our results of gold thermal conductivity using two force fields are
593 > shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
594 > these calculations,the end and middle slabs were excluded in thermal
595 > gradient regession and only used as heat source and drain in the
596 > systems. Our yielded values using EAM force field are slightly larger
597 > than those using QSC force field. However, both series are
598 > significantly smaller than experimental value by an order of more than
599 > 100. It has been verified that this difference is mainly attributed to
600 > the lack of electron interaction representation in these force field
601 > parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
602 > force field parameters in their metal thermal conductivity
603 > calculations. The Non-Equilibrium MD method they employed in their
604 > simulations produced comparable results to ours. As Zhang {\it et
605 >  al.}\cite{ISI:000231042800044} stated, thermal conductivity values
606 > are influenced mainly by force field. Therefore, it is confident to
607 > conclude that NIVS-RNEMD is applicable to metal force field system.
608  
609   \begin{figure}
610   \includegraphics[width=\linewidth]{AuGrad}
611 < \caption{Temperature gradients for crystal gold thermal conductivity.}
611 > \caption{Temperature gradients for thermal conductivity calculation of
612 >  crystal gold using QSC force field.}
613   \label{AuGrad}
614   \end{figure}
615  
# Line 555 | Line 618 | applicable to metal force field system.
618   \begin{center}
619  
620   \caption{Calculation results for thermal conductivity of crystal gold
621 <  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
622 <  calculations in parentheses. }
621 >  using QSC force field at ${\langle T\rangle}$ = 300K at various
622 >  thermal exchange rates. Errors of calculations in parentheses. }
623  
624 < \begin{tabular}{ccc}
624 > \begin{tabular}{cc}
625   \hline
626   $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
564 & This work & Previous simulations\cite{ISI:A1992HX37800010} \\
627   \hline
628 < 1.44 & 1.10(0.01) & \\
629 < 2.86 & 1.08(0.02) & \\
630 < 5.14 & 1.15(0.01) & \\
628 > 1.44 & 1.10(0.01)\\
629 > 2.86 & 1.08(0.02)\\
630 > 5.14 & 1.15(0.01)\\
631   \hline
632   \end{tabular}
633 < \label{AuThermal}
633 > \label{qscThermal}
634 > \end{center}
635 > \end{minipage}
636 > \end{table*}
637 >
638 > \begin{figure}
639 > \includegraphics[width=\linewidth]{eamGrad}
640 > \caption{Temperature gradients for thermal conductivity calculation of
641 >  crystal gold using EAM force field.}
642 > \label{eamGrad}
643 > \end{figure}
644 >
645 > \begin{table*}
646 > \begin{minipage}{\linewidth}
647 > \begin{center}
648 >
649 > \caption{Calculation results for thermal conductivity of crystal gold
650 >  using EAM force field at ${\langle T\rangle}$ = 300K at various
651 >  thermal exchange rates. Errors of calculations in parentheses. }
652 >
653 > \begin{tabular}{cc}
654 > \hline
655 > $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
656 > \hline
657 > 1.24 & 1.24(0.06)\\
658 > 2.06 & 1.37(0.04)\\
659 > 2.55 & 1.41(0.03)\\
660 > \hline
661 > \end{tabular}
662 > \label{eamThermal}
663   \end{center}
664   \end{minipage}
665   \end{table*}
666  
667 +
668   \subsection{Interfaciel Thermal Conductivity}
669 < After valid simulations of homogeneous water and gold systems using
670 < NIVS-RNEMD method, calculation of gold/water interfacial thermal
671 < conductivity was followed. It is found out that the interfacial
672 < conductance is low due to a hydrophobic surface in our system. Figure
673 < \ref{interfaceDensity} demonstrates this observance. Consequently, our
674 < reported results (Table \ref{interfaceRes}) are of two orders of
675 < magnitude smaller than our calculations on homogeneous systems.
669 > After simulations of homogeneous water and gold systems using
670 > NIVS-RNEMD method were proved valid, calculation of gold/water
671 > interfacial thermal conductivity was followed. It is found out that
672 > the low interfacial conductance is probably due to the hydrophobic
673 > surface in our system. Figure \ref{interfaceDensity} demonstrates mass
674 > density change along $z$-axis, which is perpendicular to the
675 > gold/water interface. It is observed that water density significantly
676 > decreases when approaching the surface. Under this low thermal
677 > conductance, both gold and water phase have sufficient time to
678 > eliminate temperature difference inside respectively (Figure
679 > \ref{interfaceGrad}). With indistinguishable temperature difference
680 > within respective phase, it is valid to assume that the temperature
681 > difference between gold and water on surface would be approximately
682 > the same as the difference between the gold and water phase. This
683 > assumption enables convenient calculation of $G$ using
684 > Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
685 > layer of water and gold close enough to surface, which would have
686 > greater fluctuation and lower accuracy. Reported results (Table
687 > \ref{interfaceRes}) are of two orders of magnitude smaller than our
688 > calculations on homogeneous systems, and thus have larger relative
689 > errors than our calculation results on homogeneous systems.
690  
691   \begin{figure}
692   \includegraphics[width=\linewidth]{interfaceDensity}
693   \caption{Density profile for interfacial thermal conductivity
694 <  simulation box.}
694 >  simulation box. Significant water density decrease is observed on
695 >  gold surface.}
696   \label{interfaceDensity}
697   \end{figure}
698  
699   \begin{figure}
700   \includegraphics[width=\linewidth]{interfaceGrad}
701   \caption{Temperature profiles for interfacial thermal conductivity
702 <  simulation box.}
702 >  simulation box. Temperatures of different slabs in the same phase
703 >  show no significant difference.}
704   \label{interfaceGrad}
705   \end{figure}
706  
# Line 690 | Line 798 | systems. Simulation results demonstrate its validity o
798  
799   \section{Conclusions}
800   NIVS-RNEMD simulation method is developed and tested on various
801 < systems. Simulation results demonstrate its validity of thermal
802 < conductivity calculations. NIVS-RNEMD improves non-Boltzmann-Maxwell
803 < distributions existing in previous RNEMD methods, and extends its
804 < applicability to interfacial systems. NIVS-RNEMD has also limited
805 < application on shear viscosity calculations, but under high momentum
806 < flux, it  could cause temperature difference among different
807 < dimensions. Modification is necessary to extend the applicability of
808 < NIVS-RNEMD in shear viscosity calculations.
801 > systems. Simulation results demonstrate its validity in thermal
802 > conductivity calculations, from Lennard-Jones fluid to multi-atom
803 > molecule like water and metal crystals. NIVS-RNEMD improves
804 > non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
805 > methods. Furthermore, it develops a valid means for unphysical thermal
806 > transfer between different species of molecules, and thus extends its
807 > applicability to interfacial systems. Our calculation of gold/water
808 > interfacial thermal conductivity demonstrates this advantage over
809 > previous RNEMD methods. NIVS-RNEMD has also limited application on
810 > shear viscosity calculations, but could cause temperature difference
811 > among different dimensions under high momentum flux. Modification is
812 > necessary to extend the applicability of NIVS-RNEMD in shear viscosity
813 > calculations.
814  
815   \section{Acknowledgments}
816   Support for this project was provided by the National Science
# Line 705 | Line 818 | Dame.  \newpage
818   the Center for Research Computing (CRC) at the University of Notre
819   Dame.  \newpage
820  
821 < \bibliographystyle{jcp2}
821 > \bibliographystyle{aip}
822   \bibliography{nivsRnemd}
823 +
824   \end{doublespace}
825   \end{document}
826  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines