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

Comparing stokes/stokes.tex (file contents):
Revision 3776 by skuang, Fri Dec 9 05:08:56 2011 UTC vs.
Revision 3780 by skuang, Sun Dec 11 03:22:47 2011 UTC

# Line 43 | Line 43 | Notre Dame, Indiana 46556}
43   \begin{doublespace}
44  
45   \begin{abstract}
46 <  REPLACE ABSTRACT HERE
47 <  With the Non-Isotropic Velocity Scaling (NIVS) approach to Reverse
48 <  Non-Equilibrium Molecular Dynamics (RNEMD) it is possible to impose
49 <  an unphysical thermal flux between different regions of
50 <  inhomogeneous systems such as solid / liquid interfaces.  We have
51 <  applied NIVS to compute the interfacial thermal conductance at a
52 <  metal / organic solvent interface that has been chemically capped by
53 <  butanethiol molecules.  Our calculations suggest that coupling
54 <  between the metal and liquid phases is enhanced by the capping
55 <  agents, leading to a greatly enhanced conductivity at the interface.
56 <  Specifically, the chemical bond between the metal and the capping
57 <  agent introduces a vibrational overlap that is not present without
58 <  the capping agent, and the overlap between the vibrational spectra
59 <  (metal to cap, cap to solvent) provides a mechanism for rapid
60 <  thermal transport across the interface. Our calculations also
61 <  suggest that this is a non-monotonic function of the fractional
62 <  coverage of the surface, as moderate coverages allow diffusive heat
63 <  transport of solvent molecules that have been in close contact with
64 <  the capping agent.
46 >  We present a new method for introducing stable nonequilibrium
47 >  velocity and temperature gradients in molecular dynamics simulations
48 >  of heterogeneous systems. This method conserves the linear momentum
49 >  and total energy of the system and improves previous Reverse
50 >  Non-Equilibrium Molecular Dynamics (RNEMD) methods and maintains
51 >  thermal velocity distributions. It also avoid thermal anisotropy
52 >  occured in NIVS simulations by using isotropic velocity scaling on
53 >  the molecules in specific regions of a system. To test the method,
54 >  we have computed the thermal conductivity and shear viscosity of
55 >  model liquid systems as well as the interfacial frictions of a
56 >  series of  metal/liquid interfaces.
57  
58   \end{abstract}
59  
# Line 226 | Line 218 | section will illustrate this effect.
218   diffusion of the neighboring slabs could no longer remedy this effect,
219   and nonthermal distributions would be observed. Results in later
220   section will illustrate this effect.
229 %NEW METHOD (AND NIVS) HAVE LESS PERTURBATION THAN MOMENTUM SWAPPING
221  
222   \section{Computational Details}
223   The algorithm has been implemented in our MD simulation code,
# Line 392 | Line 383 | solid-liquid interface. [MAY INCLUDE A SNAPSHOT IN FIG
383   \begin{figure}
384   \includegraphics[width=\linewidth]{defDelta}
385   \caption{The slip length $\delta$ can be obtained from a velocity
386 <  profile of a solid-liquid interface. An example of Au/hexane
387 <  interfaces is shown.}
386 >  profile of a solid-liquid interface simulation. An example of
387 >  Au/hexane interfaces is shown. Calculation for the left side is
388 >  illustrated. The right side is similar to the left side.}
389   \label{slipLength}
390   \end{figure}
391  
# Line 405 | Line 397 | data.
397   interface can be measured respectively. Further calculations and
398   characterizations of the interface can be carried out using these
399   data.
408 [MENTION IN RESULTS THAT ETA OBTAINED HERE DOES NOT NECESSARILY EQUAL
409 TO BULK VALUES]
400  
401   \section{Results and Discussions}
402   \subsection{Lennard-Jones fluid}
# Line 437 | Line 427 | calculations with various fluxes in reduced units.
427          \multicolumn{2}{c}{$\lambda^*$} &
428          \multicolumn{2}{c}{$\eta^*$} \\
429          \hline
430 <        Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
430 >        Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
431          NIVS & This work & Swapping & This work \\
432          \hline
433          0.116 & 0.16  & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
# Line 469 | Line 459 | in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j
459   expected. Furthermore, this method avoids inadvertent concomitant
460   momentum flux when only thermal flux is imposed, which could not be
461   achieved with swapping or NIVS approaches. The thermal energy exchange
462 < in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``j'')
462 > in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
463   or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
464   P^\alpha$) would not obtain this result unless thermal flux vanishes
465   (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which are trivial to apply a
# Line 498 | Line 488 | computations.
488  
489   \begin{figure}
490   \includegraphics[width=\linewidth]{tempXyz}
491 < \caption{.}
491 > \caption{Unlike the previous NIVS algorithm, the new method does not
492 >  produce a thermal anisotropy. No temperature difference between
493 >  different dimensions were observed beyond the magnitude of the error
494 >  bars. Note that the two ``hotter'' regions are caused by the shear
495 >  stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
496 >  an effect that only observed in our methods.}
497   \label{tempXyz}
498   \end{figure}
499  
# Line 525 | Line 520 | computations.
520  
521   \begin{figure}
522   \includegraphics[width=\linewidth]{velDist}
523 < \caption{.}
523 > \caption{Velocity distributions that develop under the swapping and
524 >  our methods at high flux. These distributions were obtained from
525 >  Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
526 >  swapping interval of 20 time steps). This is a relatively large flux
527 >  to demonstrate the nonthermal distributions that develop under the
528 >  swapping method. Distributions produced by our method are very close
529 >  to the ideal thermal situations.}
530   \label{vDist}
531   \end{figure}
532  
533   \subsection{Bulk SPC/E water}
534 < [WATER COMPARED TO RNEMD NIVS AND EMD]
534 > Since our method was in good performance of thermal conductivity and
535 > shear viscosity computations for simple Lennard-Jones fluid, we extend
536 > our applications of these simulations to complex fluid like SPC/E
537 > water model. A simulation cell with 1000 molecules was set up in the
538 > same manner as in \cite{kuang:164101}. For thermal conductivity
539 > simulations, measurements were taken to compare with previous RNEMD
540 > methods; for shear viscosity computations, simulations were run under
541 > a series of temperatures (with corresponding pressure relaxation using
542 > the isobaric-isothermal ensemble[CITE NIVS REF 32]), and results were
543 > compared to available data from Equilibrium MD methods[CITATIONS].
544  
545   \subsubsection{Thermal conductivity}
546 < [VSIS DOES AS WELL AS NIVS]
546 > Table \ref{spceThermal} summarizes our thermal conductivity
547 > computations under different temperatures and thermal gradients, in
548 > comparison to the previous NIVS results\cite{kuang:164101} and
549 > experimental measurements\cite{WagnerKruse}. Note that no appreciable
550 > drift of total system energy or temperature was observed when our
551 > method is applied, which indicates that our algorithm conserves total
552 > energy even for systems involving electrostatic interactions.
553  
554 < \subsubsection{Shear viscosity}
555 < [COMPARE W EMD]
554 > Measurements using our method established similar temperature
555 > gradients to the previous NIVS method. Our simulation results are in
556 > good agreement with those from previous simulations. And both methods
557 > yield values in reasonable agreement with experimental
558 > values. Simulations using moderately higher thermal gradient or those
559 > with longer gradient axis ($z$) for measurement seem to have better
560 > accuracy, from our results.
561  
562 < [MAY HAVE A FIRURE FOR DATA]
563 <
564 < [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
565 < [PUT RESULT AND FIGURE HERE IF IT WORKS]
566 < \subsection{Interfacial frictions}
567 < [SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
562 > \begin{table*}
563 >  \begin{minipage}{\linewidth}
564 >    \begin{center}
565 >      
566 >      \caption{Thermal conductivity of SPC/E water under various
567 >        imposed thermal gradients. Uncertainties are indicated in
568 >        parentheses.}
569 >      
570 >      \begin{tabular}{ccccc}
571 >        \hline\hline
572 >        $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
573 >        {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
574 >        (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
575 >        Experiment\cite{WagnerKruse} \\
576 >        \hline
577 >        300 & 0.8 & 0.815(0.027)     & 0.770(0.008) & 0.61 \\
578 >        318 & 0.8 & 0.801(0.024)     & 0.750(0.032) & 0.64 \\
579 >            & 1.6 & 0.766(0.007)     & 0.778(0.019) &      \\
580 >            & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
581 >                     twice as long.} &              &      \\
582 >        \hline\hline
583 >      \end{tabular}
584 >      \label{spceThermal}
585 >    \end{center}
586 >  \end{minipage}
587 > \end{table*}
588  
589 < qualitative agreement w interfacial thermal conductance
589 > \subsubsection{Shear viscosity}
590 > The improvement our method achieves for shear viscosity computations
591 > enables us to apply it on SPC/E water models. The series of
592 > temperatures under which our shear viscosity calculations were carried
593 > out covers the liquid range under normal pressure. Our simulations
594 > predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
595 > (Table \ref{spceShear}). Considering subtlties such as temperature or
596 > pressure/density errors in these two series of measurements, our
597 > results show no significant difference from those with EMD
598 > methods. Since each value reported using our method takes only one
599 > single trajectory of simulation, instead of average from many
600 > trajectories when using EMD, our method provides an effective means
601 > for shear viscosity computations.
602  
550 [FUTURE WORK HERE OR IN CONCLUSIONS]
551
552
603   \begin{table*}
604    \begin{minipage}{\linewidth}
605      \begin{center}
606        
607 <      \caption{Computed interfacial thermal conductance ($G$ and
608 <        $G^\prime$) values for interfaces using various models for
609 <        solvent and capping agent (or without capping agent) at
560 <        $\langle T\rangle\sim$200K.  Here ``D'' stands for deuterated
561 <        solvent or capping agent molecules. Error estimates are
562 <        indicated in parentheses.}
607 >      \caption{Computed shear viscosity of SPC/E water under different
608 >        temperatures. Results are compared to those obtained with EMD
609 >        method[CITATION]. Uncertainties are indicated in parentheses.}
610        
611 <      \begin{tabular}{llccc}
611 >      \begin{tabular}{cccc}
612          \hline\hline
613 <        Butanethiol model & Solvent & $G$ & $G^\prime$ \\
614 <        (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
613 >        $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
614 >        {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
615 >        (K) & (10$^{10}$s$^{-1}$) & This work & Previous simulations[CITATION]\\
616          \hline
617 <        UA    & UA hexane    & 131(9)    & 87(10)    \\
618 <              & UA hexane(D) & 153(5)    & 136(13)   \\
619 <              & AA hexane    & 131(6)    & 122(10)   \\
620 <              & UA toluene   & 187(16)   & 151(11)   \\
621 <              & AA toluene   & 200(36)   & 149(53)   \\
622 <        \hline
623 <        bare  & UA hexane    & 46.5(3.2) & 49.4(4.5) \\
624 <              & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
625 <              & AA hexane    & 31.0(1.4) & 29.4(1.3) \\
578 <              & UA toluene   & 70.1(1.3) & 65.8(0.5) \\
617 >        273 &  & 1.218(0.004) &  \\
618 >            &  & 1.140(0.012) &  \\
619 >        303 &  & 0.646(0.008) &  \\
620 >        318 &  & 0.536(0.007) &  \\
621 >            &  & 0.510(0.007) &  \\
622 >            &  &  &  \\
623 >        333 &  & 0.428(0.002) &  \\
624 >        363 &  & 0.279(0.014) &  \\
625 >            &  & 0.306(0.001) &  \\
626          \hline\hline
627        \end{tabular}
628 <      \label{modelTest}
628 >      \label{spceShear}
629      \end{center}
630    \end{minipage}
631   \end{table*}
632  
633 < On bare metal / solvent surfaces, different force field models for
634 < hexane yield similar results for both $G$ and $G^\prime$, and these
635 < two definitions agree with each other very well. This is primarily an
636 < indicator of weak interactions between the metal and the solvent.
633 > [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
634 > [PUT RESULTS AND FIGURE HERE IF IT WORKS]
635 > \subsection{Interfacial frictions and slip lengths}
636 > An attractive aspect of our method is the ability to apply momentum
637 > and/or thermal flux in nonhomogeneous systems, where molecules of
638 > different identities (or phases) are segregated in different
639 > regions. We have previously studied the interfacial thermal transport
640 > of a series of metal gold-liquid
641 > surfaces\cite{kuang:164101,kuang:AuThl}, and attemptions have been
642 > made to investigate the relationship between this phenomenon and the
643 > interfacial frictions.
644  
645 < For the fully-covered surfaces, the choice of force field for the
646 < capping agent and solvent has a large impact on the calculated values
647 < of $G$ and $G^\prime$.  The AA thiol to AA solvent conductivities are
648 < much larger than their UA to UA counterparts, and these values exceed
649 < the experimental estimates by a large measure.  The AA force field
650 < allows significant energy to go into C-H (or C-D) stretching modes,
651 < and since these modes are high frequency, this non-quantum behavior is
652 < likely responsible for the overestimate of the conductivity.  Compared
653 < to the AA model, the UA model yields more reasonable conductivity
654 < values with much higher computational efficiency.
645 > Table \ref{etaKappaDelta} includes these computations and previous
646 > calculations of corresponding interfacial thermal conductance. For
647 > bare Au(111) surfaces, slip boundary conditions were observed for both
648 > organic and aqueous liquid phases, corresponding to previously
649 > computed low interfacial thermal conductance. Instead, the butanethiol
650 > covered Au(111) surface appeared to be sticky to the organic liquid
651 > molecules in our simulations. We have reported conductance enhancement
652 > effect for this surface capping agent,\cite{kuang:AuThl} and these
653 > observations have a qualitative agreement with the thermal conductance
654 > results. This agreement also supports discussions on the relationship
655 > between surface wetting and slip effect and thermal conductance of the
656 > interface.[CITE BARRAT, GARDE]
657  
658 < \subsubsection{Effects due to average temperature}
658 > \begin{table*}
659 >  \begin{minipage}{\linewidth}
660 >    \begin{center}
661 >      
662 >      \caption{Computed interfacial friction coefficient values for
663 >        interfaces with various components for liquid and solid
664 >        phase. Error estimates are indicated in parentheses.}
665 >      
666 >      \begin{tabular}{llcccccc}
667 >        \hline\hline
668 >        Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
669 >        & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
670 >          \cite{kuang:164101}.} \\
671 >        surface & molecules & K & MPa & mPa$\cdot$s & Pa$\cdot$s/m & nm
672 >        & MW/m$^2$/K \\
673 >        \hline
674 >        Au(111) & hexane & 200 & 1.08 & 0.20() & 5.3$\times$10$^4$() &
675 >        3.7 & 46.5 \\
676 >                &        &     & 2.15 & 0.14() & 5.3$\times$10$^4$() &
677 >        2.7 &      \\
678 >        Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.29() & $\infty$ & 0 &
679 >        131 \\
680 >                       &        &     & 5.39 & 0.32() & $\infty$ & 0 &
681 >            \\
682 >        \hline
683 >        Au(111) & toluene & 200 & 1.08 & 0.72() & 1.?$\times$10$^5$() &
684 >        4.6 & 70.1 \\
685 >                &         &     & 2.16 & 0.54() & 1.?$\times$10$^5$() &
686 >        4.9 &      \\
687 >        Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.98() & $\infty$ & 0
688 >        & 187 \\
689 >                       &         &     & 10.8 & 0.99() & $\infty$ & 0
690 >        &     \\
691 >        \hline
692 >        Au(111) & water & 300 & 1.08 & 0.40() & 1.9$\times$10$^4$() &
693 >        20.7 & 1.65 \\
694 >                &       &     & 2.16 & 0.79() & 1.9$\times$10$^4$() &
695 >        41.9 &      \\
696 >        \hline
697 >        ice(basal) & water & 225 & 19.4 & 15.8() & $\infty$ & 0 & \\
698 >        \hline\hline
699 >      \end{tabular}
700 >      \label{etaKappaDelta}
701 >    \end{center}
702 >  \end{minipage}
703 > \end{table*}
704  
705 < We also studied the effect of average system temperature on the
706 < interfacial conductance.  The simulations are first equilibrated in
707 < the NPT ensemble at 1 atm.  The TraPPE-UA model for hexane tends to
708 < predict a lower boiling point (and liquid state density) than
709 < experiments.  This lower-density liquid phase leads to reduced contact
710 < between the hexane and butanethiol, and this accounts for our
711 < observation of lower conductance at higher temperatures.  In raising
712 < the average temperature from 200K to 250K, the density drop of
713 < $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
613 < conductance.
705 > An interesting effect alongside the surface friction change is
706 > observed on the shear viscosity of liquids in the regions close to the
707 > solid surface. Note that $\eta$ measured near a ``slip'' surface tends
708 > to be smaller than that near a ``stick'' surface. This suggests that
709 > an interface could affect the dynamic properties on its neighbor
710 > regions. It is known that diffusions of solid particles in liquid
711 > phase is affected by their surface conditions (stick or slip
712 > boundary).[CITE SCHMIDT AND SKINNER] Our observations could provide
713 > support to this phenomenon.
714  
715 < Similar behavior is observed in the TraPPE-UA model for toluene,
716 < although this model has better agreement with the experimental
717 < densities of toluene.  The expansion of the toluene liquid phase is
718 < not as significant as that of the hexane (8.3\% over 100K), and this
719 < limits the effect to $\sim$20\% drop in thermal conductivity.
715 > In addition to these previously studied interfaces, we attempt to
716 > construct ice-water interfaces and the basal plane of ice lattice was
717 > first studied. In contrast to the Au(111)/water interface, where the
718 > friction coefficient is relatively small and large slip effect
719 > presents, the ice/liquid water interface demonstrates strong
720 > interactions and appears to be sticky. The supercooled liquid phase is
721 > an order of magnitude viscous than measurements in previous
722 > section. It would be of interst to investigate the effect of different
723 > ice lattice planes (such as prism surface) on interfacial friction and
724 > corresponding liquid viscosity.
725  
621 Although we have not mapped out the behavior at a large number of
622 temperatures, is clear that there will be a strong temperature
623 dependence in the interfacial conductance when the physical properties
624 of one side of the interface (notably the density) change rapidly as a
625 function of temperature.
626
726   \section{Conclusions}
727 < [VSIS WORKS! COMBINES NICE FEATURES OF PREVIOUS RNEMD METHODS AND
728 < IMPROVEMENTS TO THEIR PROBLEMS!]
727 > Our simulations demonstrate the validity of our method in RNEMD
728 > computations of thermal conductivity and shear viscosity in atomic and
729 > molecular liquids. Our method maintains thermal velocity distributions
730 > and avoids thermal anisotropy in previous NIVS shear stress
731 > simulations, as well as retains attractive features of previous RNEMD
732 > methods. There is no {\it a priori} restrictions to the method to be
733 > applied in various ensembles, so prospective applications to
734 > extended-system methods are possible.
735  
736 < The NIVS algorithm has been applied to simulations of
737 < butanethiol-capped Au(111) surfaces in the presence of organic
738 < solvents. This algorithm allows the application of unphysical thermal
739 < flux to transfer heat between the metal and the liquid phase. With the
740 < flux applied, we were able to measure the corresponding thermal
636 < gradients and to obtain interfacial thermal conductivities. Under
637 < steady states, 2-3 ns trajectory simulations are sufficient for
638 < computation of this quantity.
736 > Furthermore, using this method, investigations can be carried out to
737 > characterize interfacial interactions. Our method is capable of
738 > effectively imposing both thermal and momentum flux accross an
739 > interface and thus facilitates studies that relates dynamic property
740 > measurements to the chemical details of an interface.
741  
742 < Our simulations have seen significant conductance enhancement in the
743 < presence of capping agent, compared with the bare gold / liquid
744 < interfaces. The vibrational coupling between the metal and the liquid
745 < phase is enhanced by a chemically-bonded capping agent. Furthermore,
746 < the coverage percentage of the capping agent plays an important role
747 < in the interfacial thermal transport process. Moderately low coverages
646 < allow higher contact between capping agent and solvent, and thus could
647 < further enhance the heat transfer process, giving a non-monotonic
648 < behavior of conductance with increasing coverage.
742 > Another attractive feature of our method is the ability of
743 > simultaneously imposing thermal and momentum flux in a
744 > system. potential researches that might be benefit include complex
745 > systems that involve thermal and momentum gradients. For example, the
746 > Soret effects under a velocity gradient would be of interest to
747 > purification and separation researches.
748  
650 Our results, particularly using the UA models, agree well with
651 available experimental data.  The AA models tend to overestimate the
652 interfacial thermal conductance in that the classically treated C-H
653 vibrations become too easily populated. Compared to the AA models, the
654 UA models have higher computational efficiency with satisfactory
655 accuracy, and thus are preferable in modeling interfacial thermal
656 transport.
657
749   \section{Acknowledgments}
750   Support for this project was provided by the National Science
751   Foundation under grant CHE-0848243. Computational time was provided by

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines