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 3778 by skuang, Sat Dec 10 02:42:55 2011 UTC vs.
Revision 3779 by skuang, Sat Dec 10 23:46:08 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 434 | Line 426 | calculations with various fluxes in reduced units.
426          \multicolumn{2}{c}{$\lambda^*$} &
427          \multicolumn{2}{c}{$\eta^*$} \\
428          \hline
429 <        Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
429 >        Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
430          NIVS & This work & Swapping & This work \\
431          \hline
432          0.116 & 0.16  & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
# Line 529 | Line 521 | computations.
521   \includegraphics[width=\linewidth]{velDist}
522   \caption{Velocity distributions that develop under the swapping and
523    our methods at high flux. These distributions were obtained from
524 <  Lennard-Jones simulations with $j_z(p_x)\sim 0.4$ (equivalent to a
524 >  Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
525    swapping interval of 20 time steps). This is a relatively large flux
526    to demonstrate the nonthermal distributions that develop under the
527    swapping method. Distributions produced by our method are very close
# Line 584 | Line 576 | accuracy, from our results.
576          300 & 0.8 & 0.815(0.027)     & 0.770(0.008) & 0.61 \\
577          318 & 0.8 & 0.801(0.024)     & 0.750(0.032) & 0.64 \\
578              & 1.6 & 0.766(0.007)     & 0.778(0.019) &      \\
579 <            & 0.8 & 0.786(0.009)$^a$ &              &      \\
579 >            & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
580 >                     twice as long.} &              &      \\
581          \hline\hline
582        \end{tabular}
590      $^a$Simulation with $L_z$ twice as long.
583        \label{spceThermal}
584      \end{center}
585    \end{minipage}
# Line 599 | Line 591 | predict a similar trend of $\eta$ vs. $T$ to EMD resul
591   temperatures under which our shear viscosity calculations were carried
592   out covers the liquid range under normal pressure. Our simulations
593   predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
594 < (Table \ref{spceShear}).
594 > (Table \ref{spceShear}). Considering subtlties such as temperature or
595 > pressure/density errors in these two series of measurements, our
596 > results show no significant difference from those with EMD
597 > methods. Since each value reported using our method takes only one
598 > single trajectory of simulation, instead of average from many
599 > trajectories when using EMD, our method provides an effective means
600 > for shear viscosity computations.
601  
602   \begin{table*}
603    \begin{minipage}{\linewidth}
# Line 633 | Line 631 | predict a similar trend of $\eta$ vs. $T$ to EMD resul
631  
632   [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
633   [PUT RESULTS AND FIGURE HERE IF IT WORKS]
634 < \subsection{Interfacial frictions}
635 < [SLIP BOUNDARY VS STICK BOUNDARY]
636 <
637 < qualitative agreement w interfacial thermal conductance
638 <
639 < [ETA OBTAINED HERE DOES NOT NECESSARILY EQUAL TO BULK VALUES]
640 <
641 <
642 < [ATTEMPT TO CONSTRUCT BASAL PLANE ICE-WATER INTERFACE]
645 <
646 < [FUTURE WORK HERE OR IN CONCLUSIONS]
634 > \subsection{Interfacial frictions and slip lengths}
635 > An attractive aspect of our method is the ability to apply momentum
636 > and/or thermal flux in nonhomogeneous systems, where molecules of
637 > different identities (or phases) are segregated in different
638 > regions. We have previously studied the interfacial thermal transport
639 > of a series of metal gold-liquid
640 > surfaces\cite{kuang:164101,kuang:AuThl}, and attemptions have been
641 > made to investigate the relationship between this phenomenon and the
642 > interfacial frictions.
643  
644 + Table \ref{etaKappaDelta} includes these computations and previous
645 + calculations of corresponding interfacial thermal conductance. For
646 + bare Au(111) surfaces, slip boundary conditions were observed for both
647 + organic and aqueous liquid phases, corresponding to previously
648 + computed low interfacial thermal conductance. Instead, the butanethiol
649 + covered Au(111) surface appeared to be sticky to the organic liquid
650 + molecules in our simulations. We have reported conductance enhancement
651 + effect for this surface capping agent,\cite{kuang:AuThl} and these
652 + observations have a qualitative agreement with the thermal conductance
653 + results. This agreement also supports discussions on the relationship
654 + between surface wetting and slip effect and thermal conductance of the
655 + interface.[CITE BARRAT, GARDE]
656  
657   \begin{table*}
658    \begin{minipage}{\linewidth}
659      \begin{center}
660        
661 <      \caption{Computed interfacial thermal conductance ($G$ and
662 <        $G^\prime$) values for interfaces using various models for
663 <        solvent and capping agent (or without capping agent) at
656 <        $\langle T\rangle\sim$200K.  Here ``D'' stands for deuterated
657 <        solvent or capping agent molecules. Error estimates are
658 <        indicated in parentheses.}
661 >      \caption{Computed interfacial friction coefficient values for
662 >        interfaces with various components for liquid and solid
663 >        phase. Error estimates are indicated in parentheses.}
664        
665 <      \begin{tabular}{llccc}
665 >      \begin{tabular}{llcccccc}
666          \hline\hline
667 <        Butanethiol model & Solvent & $G$ & $G^\prime$ \\
668 <        (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
667 >        Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
668 >        & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
669 >          \cite{kuang:164101}.} \\
670 >        surface & model & K & MPa & mPa$\cdot$s & Pa$\cdot$s/m & nm &
671 >        MW/m$^2$/K \\
672          \hline
673 <        UA    & UA hexane    & 131(9)    & 87(10)    \\
674 <              & UA hexane(D) & 153(5)    & 136(13)   \\
675 <              & AA hexane    & 131(6)    & 122(10)   \\
676 <              & UA toluene   & 187(16)   & 151(11)   \\
677 <              & AA toluene   & 200(36)   & 149(53)   \\
673 >        Au(111) & hexane & 200 & 1.08 & 0.20() & 5.3$\times$10$^4$() &
674 >        3.7 & 46.5 \\
675 >                &        &     & 2.15 & 0.14() & 5.3$\times$10$^4$() &
676 >        2.7 &      \\
677 >        Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.29() & $\infty$ & 0 &
678 >        131 \\
679 >                       &        &     & 5.39 & 0.32() & $\infty$ & 0 &
680 >            \\
681          \hline
682 <        bare  & UA hexane    & 46.5(3.2) & 49.4(4.5) \\
683 <              & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
684 <              & AA hexane    & 31.0(1.4) & 29.4(1.3) \\
685 <              & UA toluene   & 70.1(1.3) & 65.8(0.5) \\
682 >        Au(111) & toluene & 200 & 1.08 & 0.72() & 1.?$\times$10$^5$() &
683 >        4.6 & 70.1 \\
684 >                &         &     & 2.16 & 0.54() & 1.?$\times$10$^5$() &
685 >        4.9 &      \\
686 >        Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.98() & $\infty$ & 0
687 >        & 187 \\
688 >                       &         &     & 10.8 & 0.99() & $\infty$ & 0
689 >        &     \\
690 >        \hline
691 >        Au(111) & water & 300 & 1.08 & 0.40() & 1.9$\times$10$^4$() &
692 >        20.7 & 1.65 \\
693 >                &       &     & 2.16 & 0.79() & 1.9$\times$10$^4$() &
694 >        41.9 &      \\
695 >        \hline
696 >        ice(basal) & water & 225 & 19.4 & 15.8() & $\infty$ & 0 & \\
697          \hline\hline
698        \end{tabular}
699 <      \label{modelTest}
699 >      \label{etaKappaDelta}
700      \end{center}
701    \end{minipage}
702   \end{table*}
703  
704 < On bare metal / solvent surfaces, different force field models for
705 < hexane yield similar results for both $G$ and $G^\prime$, and these
706 < two definitions agree with each other very well. This is primarily an
707 < indicator of weak interactions between the metal and the solvent.
704 > An interesting effect alongside the surface friction change is
705 > observed on the shear viscosity of liquids in the regions close to the
706 > solid surface. Note that $\eta$ measured near a ``slip'' surface tends
707 > to be smaller than that near a ``stick'' surface. This suggests that
708 > an interface could affect the dynamic properties on its neighbor
709 > regions. It is known that diffusions of solid particles in liquid
710 > phase is affected by their surface conditions (stick or slip
711 > boundary).[CITE SCHMIDT AND SKINNER] Our observations could provide
712 > support to this phenomenon.
713  
714 < For the fully-covered surfaces, the choice of force field for the
715 < capping agent and solvent has a large impact on the calculated values
716 < of $G$ and $G^\prime$.  The AA thiol to AA solvent conductivities are
717 < much larger than their UA to UA counterparts, and these values exceed
718 < the experimental estimates by a large measure.  The AA force field
719 < allows significant energy to go into C-H (or C-D) stretching modes,
720 < and since these modes are high frequency, this non-quantum behavior is
721 < likely responsible for the overestimate of the conductivity.  Compared
722 < to the AA model, the UA model yields more reasonable conductivity
723 < values with much higher computational efficiency.
714 > In addition to these previously studied interfaces, we attempt to
715 > construct ice-water interfaces and the basal plane of ice lattice was
716 > first studied. In contrast to the Au(111)/water interface, where the
717 > friction coefficient is relatively small and large slip effect
718 > presents, the ice/liquid water interface demonstrates strong
719 > interactions and appears to be sticky. The supercooled liquid phase is
720 > an order of magnitude viscous than measurements in previous
721 > section. It would be of interst to investigate the effect of different
722 > ice lattice planes (such as prism surface) on interfacial friction and
723 > corresponding liquid viscosity.
724  
698 \subsubsection{Effects due to average temperature}
699
700 We also studied the effect of average system temperature on the
701 interfacial conductance.  The simulations are first equilibrated in
702 the NPT ensemble at 1 atm.  The TraPPE-UA model for hexane tends to
703 predict a lower boiling point (and liquid state density) than
704 experiments.  This lower-density liquid phase leads to reduced contact
705 between the hexane and butanethiol, and this accounts for our
706 observation of lower conductance at higher temperatures.  In raising
707 the average temperature from 200K to 250K, the density drop of
708 $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
709 conductance.
710
711 Similar behavior is observed in the TraPPE-UA model for toluene,
712 although this model has better agreement with the experimental
713 densities of toluene.  The expansion of the toluene liquid phase is
714 not as significant as that of the hexane (8.3\% over 100K), and this
715 limits the effect to $\sim$20\% drop in thermal conductivity.
716
717 Although we have not mapped out the behavior at a large number of
718 temperatures, is clear that there will be a strong temperature
719 dependence in the interfacial conductance when the physical properties
720 of one side of the interface (notably the density) change rapidly as a
721 function of temperature.
722
725   \section{Conclusions}
726 < [VSIS WORKS! COMBINES NICE FEATURES OF PREVIOUS RNEMD METHODS AND
727 < IMPROVEMENTS TO THEIR PROBLEMS! ROBUST AND VERSATILE!]
726 > Our simulations demonstrate the validity of our method in RNEMD
727 > computations of thermal conductivity and shear viscosity in atomic and
728 > molecular liquids. Our method maintains thermal velocity distributions
729 > and avoids thermal anisotropy in previous NIVS shear stress
730 > simulations, as well as retains attractive features of previous RNEMD
731 > methods. There is no {\it a priori} restrictions to the method to be
732 > applied in various ensembles, so prospective applications to
733 > extended-system methods are possible.
734  
735 < The NIVS algorithm has been applied to simulations of
736 < butanethiol-capped Au(111) surfaces in the presence of organic
737 < solvents. This algorithm allows the application of unphysical thermal
738 < flux to transfer heat between the metal and the liquid phase. With the
739 < flux applied, we were able to measure the corresponding thermal
732 < gradients and to obtain interfacial thermal conductivities. Under
733 < steady states, 2-3 ns trajectory simulations are sufficient for
734 < computation of this quantity.
735 > Furthermore, using this method, investigations can be carried out to
736 > characterize interfacial interactions. Our method is capable of
737 > effectively imposing both thermal and momentum flux accross an
738 > interface and thus facilitates studies that relates dynamic property
739 > measurements to the chemical details of an interface.
740  
741 < Our simulations have seen significant conductance enhancement in the
742 < presence of capping agent, compared with the bare gold / liquid
743 < interfaces. The vibrational coupling between the metal and the liquid
744 < phase is enhanced by a chemically-bonded capping agent. Furthermore,
745 < the coverage percentage of the capping agent plays an important role
746 < in the interfacial thermal transport process. Moderately low coverages
742 < allow higher contact between capping agent and solvent, and thus could
743 < further enhance the heat transfer process, giving a non-monotonic
744 < behavior of conductance with increasing coverage.
741 > Another attractive feature of our method is the ability of
742 > simultaneously imposing thermal and momentum flux in a
743 > system. potential researches that might be benefit include complex
744 > systems that involve thermal and momentum gradients. For example, the
745 > Soret effects under a velocity gradient would be of interest to
746 > purification and separation researches.
747  
746 Our results, particularly using the UA models, agree well with
747 available experimental data.  The AA models tend to overestimate the
748 interfacial thermal conductance in that the classically treated C-H
749 vibrations become too easily populated. Compared to the AA models, the
750 UA models have higher computational efficiency with satisfactory
751 accuracy, and thus are preferable in modeling interfacial thermal
752 transport.
753
748   \section{Acknowledgments}
749   Support for this project was provided by the National Science
750   Foundation under grant CHE-0848243. Computational time was provided by

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines