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 3579 by skuang, Tue Mar 23 22:46:17 2010 UTC vs.
Revision 3595 by skuang, Mon Apr 19 21:54:15 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
104 > methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
105 > because it imposes what is typically difficult to measure (a flux or
106 > stress) and it is typically much easier to compute momentum gradients
107 > or strains (the response).  For similar reasons, RNEMD is also
108 > preferable to slowly-converging equilibrium methods for measuring
109 > thermal conductivity and shear viscosity (using Green-Kubo
110 > relations\cite{daivis:541,mondello:9327} or the Helfand moment
111 > approach of Viscardy {\it et
112    al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
113   computing difficult to measure quantities.
114  
# Line 100 | Line 118 | Recently, Tenney and Maginn\cite{ISI:000273472300004}
118   typically samples from the same manifold of states in the
119   microcanonical ensemble.
120  
121 < Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
121 > Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
122   some problems with the original RNEMD swap technique.  Notably, large
123   momentum fluxes (equivalent to frequent momentum swaps between the
124   slabs) can result in ``notched'', ``peaked'' and generally non-thermal
# Line 114 | Line 132 | develop the method for determining the scaling constra
132   (conservation of linear momentum and total energy, compatibility with
133   periodic boundary conditions) while establishing true thermal
134   distributions in each of the two slabs.  In the next section, we
135 < develop the method for determining the scaling constraints.  We then
135 > present the method for determining the scaling constraints.  We then
136   test the method on both single component, multi-component, and
137   non-isotropic mixtures and show that it is capable of providing
138   reasonable estimates of the thermal conductivity and shear viscosity
139   in these cases.
140  
141   \section{Methodology}
142 < We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
143 < system is partitioned into a series of thin slabs along a particular
142 > We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
143 > periodic system is partitioned into a series of thin slabs along one
144   axis ($z$).  One of the slabs at the end of the periodic box is
145   designated the ``hot'' slab, while the slab in the center of the box
146   is designated the ``cold'' slab.  The artificial momentum flux will be
# Line 130 | Line 148 | moves.  For molecules $\{i\}$ located within the cold
148   hot slab.
149  
150   Rather than using momentum swaps, we use a series of velocity scaling
151 < moves.  For molecules $\{i\}$ located within the cold slab,
151 > moves.  For molecules $\{i\}$  located within the cold slab,
152   \begin{equation}
153   \vec{v}_i \leftarrow \left( \begin{array}{ccc}
154   x & 0 & 0 \\
# Line 150 | Line 168 | Conservation of linear momentum in each of the three d
168   \end{equation}
169  
170   Conservation of linear momentum in each of the three directions
171 < ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
171 > ($\alpha = x,y,z$) ties the values of the hot and cold scaling
172   parameters together:
173   \begin{equation}
174   P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
# Line 184 | Line 202 | Substituting in the expressions for the hot scaling pa
202   similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
203   Substituting in the expressions for the hot scaling parameters
204   ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
205 < {\it constraint ellipsoid equation}:
205 > {\it constraint ellipsoid}:
206   \begin{equation}
207   \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
208   \label{eq:constraintEllipsoid}
# Line 198 | Line 216 | This ellipsoid equation defines the set of cold slab s
216   c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
217   \label{eq:constraintEllipsoidConsts}
218   \end{eqnarray}
219 < This ellipsoid equation defines the set of cold slab scaling
220 < parameters which can be applied while preserving both linear momentum
221 < in all three directions as well as kinetic energy.
219 > This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
220 > cold slab scaling parameters which can be applied while preserving
221 > both linear momentum in all three directions as well as total kinetic
222 > energy.
223  
224   The goal of using velocity scaling variables is to transfer linear
225   momentum or kinetic energy from the cold slab to the hot slab.  If the
# Line 215 | Line 234 | The spatial extent of the {\it thermal flux ellipsoid
234   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
235   \label{eq:fluxEllipsoid}
236   \end{equation}
237 < The spatial extent of the {\it thermal flux ellipsoid equation} is
238 < governed both by a targetted value, $J_z$ as well as the instantaneous
239 < values of the kinetic energy components in the cold bin.
237 > The spatial extent of the {\it thermal flux ellipsoid} is governed
238 > both by a targetted value, $J_z$ as well as the instantaneous values
239 > of the kinetic energy components in the cold bin.
240  
241   To satisfy an energetic flux as well as the conservation constraints,
242 < it is sufficient to determine the points ${x,y,z}$ which lie on both
243 < the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
244 < flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
245 < the two ellipsoids in 3-dimensional space.
242 > we must determine the points ${x,y,z}$ which lie on both the
243 > constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
244 > ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
245 > two ellipsoids in 3-dimensional space.
246  
247   \begin{figure}
248   \includegraphics[width=\linewidth]{ellipsoids}
249 < \caption{Scaling points which maintain both constant energy and
250 <  constant linear momentum of the system lie on the surface of the
251 <  {\it constraint ellipsoid} while points which generate the target
252 <  momentum flux lie on the surface of the {\it flux ellipsoid}.  The
253 <  velocity distributions in the cold bin are scaled by only those
254 <  points which lie on both ellipsoids.}
249 > \caption{Illustration from the perspective of a space having cold
250 >  slab scaling coefficients as its coordinates. Scaling points which
251 >  maintain both constant energy and constant linear momentum of the
252 >  system lie on the surface of the {\it constraint ellipsoid} while
253 >  points which generate the target momentum flux lie on the surface of
254 >  the {\it flux ellipsoid}. The velocity distributions in the cold bin
255 >  are scaled by only those points which lie on both ellipsoids.}
256   \label{ellipsoids}
257   \end{figure}
258  
259 < One may also define momentum flux (say along the $x$-direction) as:
259 > One may also define {\it momentum} flux (say along the $x$-direction) as:
260   \begin{equation}
261   (1-x) P_c^x = j_z(p_x)\Delta t
262   \label{eq:fluxPlane}
263   \end{equation}
264 < The above {\it momentum flux equation} is essentially a plane which is
265 < perpendicular to the $x$-axis, with its position governed both by a
266 < target value, $j_z(p_x)$ as well as the instantaneous value of the
247 < momentum along the $x$-direction.
264 > The above {\it momentum flux plane} is perpendicular to the $x$-axis,
265 > with its position governed both by a target value, $j_z(p_x)$ as well
266 > as the instantaneous value of the momentum along the $x$-direction.
267  
268 < Similarly, to satisfy a momentum flux as well as the conservation
269 < constraints, it is sufficient to determine the points ${x,y,z}$ which
270 < lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
271 < and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
272 < an ellipsoid and a plane in 3-dimensional space.
268 > In order to satisfy a momentum flux as well as the conservation
269 > constraints, we must determine the points ${x,y,z}$ which lie on both
270 > the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
271 > flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
272 > ellipsoid and a plane in 3-dimensional space.  
273  
274 < To summarize, by solving respective equation sets, one can determine
275 < possible sets of scaling variables for cold slab. And corresponding
276 < sets of scaling variables for hot slab can be determine as well.
274 > In both the momentum and energy flux scenarios, valid scaling
275 > parameters are arrived at by solving geometric intersection problems
276 > in $x, y, z$ space in order to obtain cold slab scaling parameters.
277 > Once the scaling variables for the cold slab are known, the hot slab
278 > scaling has also been determined.
279  
280 +
281   The following problem will be choosing an optimal set of scaling
282   variables among the possible sets. Although this method is inherently
283   non-isotropic, the goal is still to maintain the system as isotropic
# Line 263 | Line 285 | large perturbation to the system. Therefore, one appro
285   energies in different directions could become as close as each other
286   after each scaling. Simultaneously, one would also like each scaling
287   as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
288 < large perturbation to the system. Therefore, one approach to obtain the
289 < scaling variables would be constructing an criteria function, with
288 > large perturbation to the system. Therefore, one approach to obtain
289 > 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.
292  
# Line 314 | Line 336 | to Tenney {\it et al.}\cite{ISI:000273472300004}, a se
336   periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
337   the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
338   most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
339 < to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
339 > to Tenney {\it et al.}\cite{Maginn:2010}, a series of swapping
340   frequency were chosen. According to each result from swapping
341   RNEMD, scaling RNEMD simulations were run with the target momentum
342   flux set to produce a similar momentum flux, and consequently shear
# Line 375 | Line 397 | liquid water (Extended Simple Point Charge model) and
397   Another series of our simulation is the calculation of interfacial
398   thermal conductivity of a Au/H$_2$O system. Respective calculations of
399   liquid water (Extended Simple Point Charge model) and crystal gold
400 < (Quantum Sutton-Chen potential) thermal conductivity were performed
401 < and compared with current results to ensure the validity of
402 < NIVS-RNEMD. After that, a mixture system was simulated.
400 > thermal conductivity were performed and compared with current results
401 > to ensure the validity of NIVS-RNEMD. After that, a mixture system was
402 > simulated.
403  
404   For thermal conductivity calculation of bulk water, a simulation box
405   consisting of 1000 molecules were first equilibrated under ambient
# Line 389 | Line 411 | protocol. The face-centered cubic crystal simulation b
411   process was similar to Lennard-Jones fluid system.
412  
413   Thermal conductivity calculation of bulk crystal gold used a similar
414 < protocol. The face-centered cubic crystal simulation box consists of
414 > protocol. Two types of force field parameters, Embedded Atom Method
415 > (EAM) and Quantum Sutten-Chen (QSC) force field were used
416 > respectively. The face-centered cubic crystal simulation box consists of
417   2880 Au atoms. The lattice was first allowed volume change to relax
418   under ambient temperature and pressure. Equilibrations in canonical and
419   microcanonical ensemble were followed in order. With the simulation
# Line 458 | Line 482 | $\lambda^*_{scale}$\\
482  
483   \begin{figure}
484   \includegraphics[width=\linewidth]{thermalGrad}
485 < \caption{Temperature gradients under various kinetic energy flux of
486 <  thermal conductivity simulations}
485 > \caption{NIVS-RNEMD method introduced similar temperature gradients
486 >  compared to ``swapping'' method under various kinetic energy flux in
487 >  thermal conductivity simulations.}
488   \label{thermalGrad}
489   \end{figure}
490  
# Line 470 | Line 495 | could deviate from Maxwell-Boltzmann distribution. Fig
495   that under relatively high unphysical kinetic energy flux, speed and
496   velocity distribution of molecules in slabs where swapping occured
497   could deviate from Maxwell-Boltzmann distribution. Figure
498 < \ref{histSwap} illustrates how these distributions deviate from an
498 > \ref{thermalHist} a) illustrates how these distributions deviate from an
499   ideal distribution. In high temperature slab, probability density in
500   low speed is confidently smaller than ideal curve fit; in low
501   temperature slab, probability density in high speed is smaller than
# Line 491 | Line 516 | choice of swapping rate to satisfy these above restric
516   interference. Consequently, swapping RNEMD has a relatively narrow
517   choice of swapping rate to satisfy these above restrictions.
518  
494 \begin{figure}
495 \includegraphics[width=\linewidth]{histSwap}
496 \caption{Speed distribution for thermal conductivity using swapping
497  RNEMD. Shown is from the simulation with 250 fs exchange interval.}
498 \label{histSwap}
499 \end{figure}
500
519   Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
520 < curve fit (Figure \ref{histScale}). Essentially, after scaling, a
520 > curve fit (Figure \ref{thermalHist} b). Essentially, after scaling, a
521   Gaussian distribution function would remain Gaussian. Although a
522   single scaling is non-isotropic in all three dimensions, our scaling
523   coefficient criteria could help maintian the scaling region as
# Line 512 | Line 530 | to the distribution of velocity and speed in the excha
530   to the distribution of velocity and speed in the exchange regions.
531  
532   \begin{figure}
533 < \includegraphics[width=\linewidth]{histScale}
534 < \caption{Speed distribution for thermal conductivity using scaling
535 <  RNEMD. Shown is from the simulation with an equilvalent thermal flux
536 <  as an 250 fs exchange interval swapping simulation.}
537 < \label{histScale}
533 > \includegraphics[width=\linewidth]{thermalHist}
534 > \caption{Speed distribution for thermal conductivity using a)
535 >  ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
536 >  simulations with an exchange or equilvalent exchange interval of 250
537 >  fs. In circled areas, distributions from ``swapping'' RNEMD
538 >  simulation have deviation from ideal Maxwell-Boltzmann distribution
539 >  (curves fit for each distribution).}
540 > \label{thermalHist}
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
545 > Bedrov {\it et al.}\cite{Bedrov:2000}, which employed the
546   previous swapping RNEMD method for their calculation. Bedrov {\it et
547 <  al.}\cite{ISI:000090151400044} argued that exchange of the molecule
547 >  al.}\cite{Bedrov:2000} 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
# Line 539 | Line 560 | multi-atom molecular system.
560  
561   \begin{figure}
562   \includegraphics[width=\linewidth]{spceGrad}
563 < \caption{Temperature gradients for SPC/E water thermal conductivity.}
563 > \caption{Temperature gradients in SPC/E water thermal conductivity
564 >  simulations.}
565   \label{spceGrad}
566   \end{figure}
567  
# Line 554 | Line 576 | $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \
576   \begin{tabular}{cccc}
577   \hline
578   $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
579 < & This work & Previous simulations\cite{ISI:000090151400044} &
579 > & This work & Previous simulations\cite{Bedrov:2000} &
580   Experiment$^a$\\
581   \hline
582   0.38 & 0.816(0.044) & & 0.64\\
# Line 568 | Line 590 | Our results of gold thermal conductivity using QSC for
590   \end{table*}
591  
592   \subsubsection{Crystal Gold}
593 < Our results of gold thermal conductivity using QSC force field are
594 < shown in Table \ref{AuThermal}. Although our calculation is smaller
595 < than experimental value by an order of more than 100, this difference
596 < is mainly attributed to the lack of electron interaction
597 < representation in our force field parameters. Richardson {\it et
598 <  al.}\cite{ISI:A1992HX37800010} using similar force field parameters
599 < in their metal thermal conductivity calculations. The EMD method they
600 < employed in their simulations produced comparable results to
601 < ours. Therefore, it is confident to conclude that NIVS-RNEMD is
602 < applicable to metal force field system.
593 > Our results of gold thermal conductivity using two force fields are
594 > shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
595 > these calculations,the end and middle slabs were excluded in thermal
596 > gradient regession and only used as heat source and drain in the
597 > systems. Our yielded values using EAM force field are slightly larger
598 > than those using QSC force field. However, both series are
599 > significantly smaller than experimental value by an order of more than
600 > 100. It has been verified that this difference is mainly attributed to
601 > the lack of electron interaction representation in these force field
602 > parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
603 > force field parameters in their metal thermal conductivity
604 > calculations. The Non-Equilibrium MD method they employed in their
605 > simulations produced comparable results to ours. As Zhang {\it et
606 >  al.}\cite{ISI:000231042800044} stated, thermal conductivity values
607 > are influenced mainly by force field. Therefore, it is confident to
608 > conclude that NIVS-RNEMD is applicable to metal force field system.
609  
610   \begin{figure}
611   \includegraphics[width=\linewidth]{AuGrad}
612 < \caption{Temperature gradients for crystal gold thermal conductivity.}
612 > \caption{Temperature gradients for thermal conductivity calculation of
613 >  crystal gold using QSC force field.}
614   \label{AuGrad}
615   \end{figure}
616  
# Line 590 | Line 619 | applicable to metal force field system.
619   \begin{center}
620  
621   \caption{Calculation results for thermal conductivity of crystal gold
622 <  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
623 <  calculations in parentheses. }
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
# Line 602 | Line 631 | $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
631   5.14 & 1.15(0.01)\\
632   \hline
633   \end{tabular}
634 < \label{AuThermal}
634 > \label{qscThermal}
635   \end{center}
636   \end{minipage}
637   \end{table*}
638  
610 \subsection{Interfaciel Thermal Conductivity}
611 After valid simulations of homogeneous water and gold systems using
612 NIVS-RNEMD method, calculation of gold/water interfacial thermal
613 conductivity was followed. It is found out that the interfacial
614 conductance is low due to a hydrophobic surface in our system. Figure
615 \ref{interfaceDensity} demonstrates this observance. Consequently, our
616 reported results (Table \ref{interfaceRes}) are of two orders of
617 magnitude smaller than our calculations on homogeneous systems.
618
639   \begin{figure}
640 < \includegraphics[width=\linewidth]{interfaceDensity}
641 < \caption{Density profile for interfacial thermal conductivity
642 <  simulation box.}
643 < \label{interfaceDensity}
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 +
646 + \begin{table*}
647 + \begin{minipage}{\linewidth}
648 + \begin{center}
649 +
650 + \caption{Calculation results for thermal conductivity of crystal gold
651 +  using EAM force field at ${\langle T\rangle}$ = 300K at various
652 +  thermal exchange rates. Errors of calculations in parentheses. }
653 +
654 + \begin{tabular}{cc}
655 + \hline
656 + $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
657 + \hline
658 + 1.24 & 1.24(0.06)\\
659 + 2.06 & 1.37(0.04)\\
660 + 2.55 & 1.41(0.03)\\
661 + \hline
662 + \end{tabular}
663 + \label{eamThermal}
664 + \end{center}
665 + \end{minipage}
666 + \end{table*}
667  
668 +
669 + \subsection{Interfaciel Thermal Conductivity}
670 + After simulations of homogeneous water and gold systems using
671 + NIVS-RNEMD method were proved valid, calculation of gold/water
672 + interfacial thermal conductivity was followed. It is found out that
673 + the low interfacial conductance is probably due to the hydrophobic
674 + surface in our system. Figure \ref{interface} (a) demonstrates mass
675 + density change along $z$-axis, which is perpendicular to the
676 + gold/water interface. It is observed that water density significantly
677 + decreases when approaching the surface. Under this low thermal
678 + conductance, both gold and water phase have sufficient time to
679 + eliminate temperature difference inside respectively (Figure
680 + \ref{interface} b). With indistinguishable temperature difference
681 + within respective phase, it is valid to assume that the temperature
682 + difference between gold and water on surface would be approximately
683 + the same as the difference between the gold and water phase. This
684 + assumption enables convenient calculation of $G$ using
685 + Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
686 + layer of water and gold close enough to surface, which would have
687 + greater fluctuation and lower accuracy. Reported results (Table
688 + \ref{interfaceRes}) are of two orders of magnitude smaller than our
689 + calculations on homogeneous systems, and thus have larger relative
690 + errors than our calculation results on homogeneous systems.
691 +
692   \begin{figure}
693 < \includegraphics[width=\linewidth]{interfaceGrad}
694 < \caption{Temperature profiles for interfacial thermal conductivity
695 <  simulation box.}
696 < \label{interfaceGrad}
693 > \includegraphics[width=\linewidth]{interface}
694 > \caption{Simulation results for Gold/Water interfacial thermal
695 >  conductivity: (a) Significant water density decrease is observed on
696 >  crystalline gold surface. (b) Temperature profiles for a series of
697 >  simulations. Temperatures of different slabs in the same phase show
698 >  no significant differences.}
699 > \label{interface}
700   \end{figure}
701  
702   \begin{table*}
# Line 663 | Line 732 | are shown in Figure \ref{shearGrad}.
732   momentum flux would theoretically result in swap method. All the scale
733   method results were from simulations that had a scaling interval of 10
734   time steps. The average molecular momentum gradients of these samples
735 < are shown in Figure \ref{shearGrad}.
735 > are shown in Figure \ref{shear} (a) and (b).
736  
737   \begin{table*}
738   \begin{minipage}{\linewidth}
# Line 690 | Line 759 | Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
759   \end{table*}
760  
761   \begin{figure}
762 < \includegraphics[width=\linewidth]{shearGrad}
763 < \caption{Average momentum gradients of shear viscosity simulations}
764 < \label{shearGrad}
762 > \includegraphics[width=\linewidth]{shear}
763 > \caption{Average momentum gradients in shear viscosity simulations,
764 >  using (a) ``swapping'' method and (b) NIVS-RNEMD method
765 >  respectively. (c) Temperature difference among x and y, z dimensions
766 >  observed when using NIVS-RNEMD with equivalent exchange interval of
767 >  500 fs.}
768 > \label{shear}
769   \end{figure}
770  
698 \begin{figure}
699 \includegraphics[width=\linewidth]{shearTempScale}
700 \caption{Temperature profile for scaling RNEMD simulation.}
701 \label{shearTempScale}
702 \end{figure}
771   However, observations of temperatures along three dimensions show that
772   inhomogeneity occurs in scaling RNEMD simulations, particularly in the
773 < two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
773 > two slabs which were scaled. Figure \ref{shear} (c) indicate that with
774   relatively large imposed momentum flux, the temperature difference among $x$
775   and the other two dimensions was significant. This would result from the
776   algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
# Line 724 | Line 792 | systems. Simulation results demonstrate its validity o
792  
793   \section{Conclusions}
794   NIVS-RNEMD simulation method is developed and tested on various
795 < systems. Simulation results demonstrate its validity of thermal
796 < conductivity calculations. NIVS-RNEMD improves non-Boltzmann-Maxwell
797 < distributions existing in previous RNEMD methods, and extends its
798 < applicability to interfacial systems. NIVS-RNEMD has also limited
799 < application on shear viscosity calculations, but under high momentum
800 < flux, it  could cause temperature difference among different
801 < dimensions. Modification is necessary to extend the applicability of
802 < NIVS-RNEMD in shear viscosity calculations.
795 > systems. Simulation results demonstrate its validity in thermal
796 > conductivity calculations, from Lennard-Jones fluid to multi-atom
797 > molecule like water and metal crystals. NIVS-RNEMD improves
798 > non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
799 > methods. Furthermore, it develops a valid means for unphysical thermal
800 > transfer between different species of molecules, and thus extends its
801 > applicability to interfacial systems. Our calculation of gold/water
802 > interfacial thermal conductivity demonstrates this advantage over
803 > previous RNEMD methods. NIVS-RNEMD has also limited application on
804 > shear viscosity calculations, but could cause temperature difference
805 > among different dimensions under high momentum flux. Modification is
806 > necessary to extend the applicability of NIVS-RNEMD in shear viscosity
807 > calculations.
808  
809   \section{Acknowledgments}
810   Support for this project was provided by the National Science
# Line 739 | Line 812 | Dame.  \newpage
812   the Center for Research Computing (CRC) at the University of Notre
813   Dame.  \newpage
814  
815 < \bibliographystyle{jcp2}
815 > \bibliographystyle{aip}
816   \bibliography{nivsRnemd}
817 +
818   \end{doublespace}
819   \end{document}
820  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines