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 3775 by skuang, Thu Dec 8 01:34:21 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 218 | Line 210 | thermal anisotropy when applying a momentum flux.
210   scaling. More importantly, separating the momentum flux imposing from
211   velocity scaling avoids the underlying cause that NIVS produced
212   thermal anisotropy when applying a momentum flux.
221 %NEW METHOD DOESN'T CAUSE UNDESIRED CONCOMITENT MOMENTUM FLUX WHEN
222 %IMPOSING A THERMAL FLUX
213  
214   The advantages of the approach over the original momentum swapping
215   approach lies in its nature to preserve a Gaussian
# Line 228 | 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.
231 %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 389 | Line 378 | solid-liquid interface.
378   so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
379   and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
380   illustrates how this quantity is defined and computed for a
381 < solid-liquid interface.
381 > solid-liquid interface. [MAY INCLUDE A SNAPSHOT IN FIGURE]
382  
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 407 | 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.
410 [MENTION IN RESULTS THAT ETA OBTAINED HERE DOES NOT NECESSARILY EQUAL
411 TO BULK VALUES]
400  
401 < \section{Results}
402 < [L-J COMPARED TO RNEMD NIVS; WATER COMPARED TO RNEMD NIVS AND EMD;
403 < SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
404 <
405 < There are many factors contributing to the measured interfacial
406 < conductance; some of these factors are physically motivated
407 < (e.g. coverage of the surface by the capping agent coverage and
408 < solvent identity), while some are governed by parameters of the
409 < methodology (e.g. applied flux and the formulas used to obtain the
410 < conductance). In this section we discuss the major physical and
411 < calculational effects on the computed conductivity.
401 > \section{Results and Discussions}
402 > \subsection{Lennard-Jones fluid}
403 > Our orthorhombic simulation cell of Lennard-Jones fluid has identical
404 > parameters to our previous work\cite{kuang:164101} to facilitate
405 > comparison. Thermal conductivitis and shear viscosities were computed
406 > with the algorithm applied to the simulations. The results of thermal
407 > conductivity are compared with our previous NIVS algorithm. However,
408 > since the NIVS algorithm could produce temperature anisotropy for
409 > shear viscocity computations, these results are instead compared to
410 > the momentum swapping approaches. Table \ref{LJ} lists these
411 > calculations with various fluxes in reduced units.
412  
413 < \subsection{Effects due to capping agent coverage}
413 > \begin{table*}
414 >  \begin{minipage}{\linewidth}
415 >    \begin{center}
416  
417 < A series of different initial conditions with a range of surface
418 < coverages was prepared and solvated with various with both of the
419 < solvent molecules. These systems were then equilibrated and their
420 < interfacial thermal conductivity was measured with the NIVS
421 < algorithm. Figure \ref{coverage} demonstrates the trend of conductance
422 < with respect to surface coverage.
417 >      \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
418 >        ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
419 >        ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
420 >        at various momentum fluxes.  The new method yields similar
421 >        results to previous RNEMD methods. All results are reported in
422 >        reduced unit. Uncertainties are indicated in parentheses.}
423 >      
424 >      \begin{tabular}{cccccc}
425 >        \hline\hline
426 >        \multicolumn{2}{c}{Momentum Exchange} &
427 >        \multicolumn{2}{c}{$\lambda^*$} &
428 >        \multicolumn{2}{c}{$\eta^*$} \\
429 >        \hline
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)\\
434 >        0.232 & 0.09  & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
435 >        0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
436 >        0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
437 >        1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
438 >        \hline\hline
439 >      \end{tabular}
440 >      \label{LJ}
441 >    \end{center}
442 >  \end{minipage}
443 > \end{table*}
444 >
445 > \subsubsection{Thermal conductivity}
446 > Our thermal conductivity calculations with this method yields
447 > comparable results to the previous NIVS algorithm. This indicates that
448 > the thermal gradients rendered using this method are also close to
449 > previous RNEMD methods. Simulations with moderately higher thermal
450 > fluxes tend to yield more reliable thermal gradients and thus avoid
451 > large errors, while overly high thermal fluxes could introduce side
452 > effects such as non-linear temperature gradient response or
453 > inadvertent phase transitions.
454  
455 < \begin{figure}
456 < \includegraphics[width=\linewidth]{coverage}
457 < \caption{The interfacial thermal conductivity ($G$) has a
458 <  non-monotonic dependence on the degree of surface capping.  This
459 <  data is for the Au(111) / butanethiol / solvent interface with
460 <  various UA force fields at $\langle T\rangle \sim $200K.}
461 < \label{coverage}
462 < \end{figure}
455 > Since the scaling operation is isotropic in this method, one does not
456 > need extra care to ensure temperature isotropy between the $x$, $y$
457 > and $z$ axes, while thermal anisotropy might happen if the criteria
458 > function for choosing scaling coefficients does not perform as
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 ``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
466 > thermal flux). In this sense, this method contributes to having
467 > minimal perturbation to a simulation while imposing thermal flux.
468  
469 < In partially covered surfaces, the derivative definition for
470 < $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
471 < location of maximum change of $\lambda$ becomes washed out.  The
472 < discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
473 < Gibbs dividing surface is still well-defined. Therefore, $G$ (not
474 < $G^\prime$) was used in this section.
469 > \subsubsection{Shear viscosity}
470 > Table \ref{LJ} also compares our shear viscosity results with momentum
471 > swapping approach. Our calculations show that our method predicted
472 > similar values for shear viscosities to the momentum swapping
473 > approach, as well as the velocity gradient profiles. Moderately larger
474 > momentum fluxes are helpful to reduce the errors of measured velocity
475 > gradients and thus the final result. However, it is pointed out that
476 > the momentum swapping approach tends to produce nonthermal velocity
477 > distributions.\cite{Maginn:2010}
478  
479 < From Figure \ref{coverage}, one can see the significance of the
480 < presence of capping agents. When even a small fraction of the Au(111)
481 < surface sites are covered with butanethiols, the conductivity exhibits
482 < an enhancement by at least a factor of 3.  Capping agents are clearly
483 < playing a major role in thermal transport at metal / organic solvent
484 < surfaces.
479 > To examine that temperature isotropy holds in simulations using our
480 > method, we measured the three one-dimensional temperatures in each of
481 > the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
482 > temperatures were calculated after subtracting the effects from bulk
483 > velocities of the slabs. The one-dimensional temperature profiles
484 > showed no observable difference between the three dimensions. This
485 > ensures that isotropic scaling automatically preserves temperature
486 > isotropy and that our method is useful in shear viscosity
487 > computations.
488  
489 < We note a non-monotonic behavior in the interfacial conductance as a
490 < function of surface coverage. The maximum conductance (largest $G$)
491 < happens when the surfaces are about 75\% covered with butanethiol
492 < caps.  The reason for this behavior is not entirely clear.  One
493 < explanation is that incomplete butanethiol coverage allows small gaps
494 < between butanethiols to form. These gaps can be filled by transient
495 < solvent molecules.  These solvent molecules couple very strongly with
496 < the hot capping agent molecules near the surface, and can then carry
497 < away (diffusively) the excess thermal energy from the surface.
489 > \begin{figure}
490 > \includegraphics[width=\linewidth]{tempXyz}
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  
500 < There appears to be a competition between the conduction of the
501 < thermal energy away from the surface by the capping agents (enhanced
502 < by greater coverage) and the coupling of the capping agents with the
503 < solvent (enhanced by interdigitation at lower coverages).  This
504 < competition would lead to the non-monotonic coverage behavior observed
505 < here.
506 <
507 < Results for rigid body toluene solvent, as well as the UA hexane, are
508 < within the ranges expected from prior experimental
509 < work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
510 < that explicit hydrogen atoms might not be required for modeling
511 < thermal transport in these systems.  C-H vibrational modes do not see
512 < significant excited state population at low temperatures, and are not
513 < likely to carry lower frequency excitations from the solid layer into
514 < the bulk liquid.
500 > Furthermore, the velocity distribution profiles are tested by imposing
501 > a large shear stress into the simulations. Figure \ref{vDist}
502 > demonstrates how our method is able to maintain thermal velocity
503 > distributions against the momentum swapping approach even under large
504 > imposed fluxes. Previous swapping methods tend to deplete particles of
505 > positive velocities in the negative velocity slab (``c'') and vice
506 > versa in slab ``h'', where the distributions leave a notch. This
507 > problematic profiles become significant when the imposed-flux becomes
508 > larger and diffusions from neighboring slabs could not offset the
509 > depletion. Simutaneously, abnormal peaks appear corresponding to
510 > excessive velocity swapped from the other slab. This nonthermal
511 > distributions limit applications of the swapping approach in shear
512 > stress simulations. Our method avoids the above problematic
513 > distributions by altering the means of applying momentum
514 > fluxes. Comparatively, velocity distributions recorded from
515 > simulations with our method is so close to the ideal thermal
516 > prediction that no observable difference is shown in Figure
517 > \ref{vDist}. Conclusively, our method avoids problems happened in
518 > previous RNEMD methods and provides a useful means for shear viscosity
519 > computations.
520  
521 < The toluene solvent does not exhibit the same behavior as hexane in
522 < that $G$ remains at approximately the same magnitude when the capping
523 < coverage increases from 25\% to 75\%.  Toluene, as a rigid planar
524 < molecule, cannot occupy the relatively small gaps between the capping
525 < agents as easily as the chain-like {\it n}-hexane.  The effect of
526 < solvent coupling to the capping agent is therefore weaker in toluene
527 < except at the very lowest coverage levels.  This effect counters the
528 < coverage-dependent conduction of heat away from the metal surface,
529 < leading to a much flatter $G$ vs. coverage trend than is observed in
530 < {\it n}-hexane.
521 > \begin{figure}
522 > \includegraphics[width=\linewidth]{velDist}
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{Effects due to Solvent \& Solvent Models}
534 < In addition to UA solvent and capping agent models, AA models have
535 < also been included in our simulations.  In most of this work, the same
536 < (UA or AA) model for solvent and capping agent was used, but it is
537 < also possible to utilize different models for different components.
538 < We have also included isotopic substitutions (Hydrogen to Deuterium)
539 < to decrease the explicit vibrational overlap between solvent and
540 < capping agent. Table \ref{modelTest} summarizes the results of these
541 < studies.
533 > \subsection{Bulk SPC/E water}
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 + 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 + 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   \begin{table*}
563    \begin{minipage}{\linewidth}
564      \begin{center}
565        
566 <      \caption{Computed interfacial thermal conductance ($G$ and
567 <        $G^\prime$) values for interfaces using various models for
568 <        solvent and capping agent (or without capping agent) at
511 <        $\langle T\rangle\sim$200K.  Here ``D'' stands for deuterated
512 <        solvent or capping agent molecules. Error estimates are
513 <        indicated in parentheses.}
566 >      \caption{Thermal conductivity of SPC/E water under various
567 >        imposed thermal gradients. Uncertainties are indicated in
568 >        parentheses.}
569        
570 <      \begin{tabular}{llccc}
570 >      \begin{tabular}{ccccc}
571          \hline\hline
572 <        Butanethiol model & Solvent & $G$ & $G^\prime$ \\
573 <        (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
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 <        UA    & UA hexane    & 131(9)    & 87(10)    \\
578 <              & UA hexane(D) & 153(5)    & 136(13)   \\
579 <              & AA hexane    & 131(6)    & 122(10)   \\
580 <              & UA toluene   & 187(16)   & 151(11)   \\
581 <              & AA toluene   & 200(36)   & 149(53)   \\
525 <        \hline
526 <        AA    & UA hexane    & 116(9)    & 129(8)    \\
527 <              & AA hexane    & 442(14)   & 356(31)   \\
528 <              & AA hexane(D) & 222(12)   & 234(54)   \\
529 <              & UA toluene   & 125(25)   & 97(60)    \\
530 <              & AA toluene   & 487(56)   & 290(42)   \\
531 <        \hline
532 <        AA(D) & UA hexane    & 158(25)   & 172(4)    \\
533 <              & AA hexane    & 243(29)   & 191(11)   \\
534 <              & AA toluene   & 364(36)   & 322(67)   \\
535 <        \hline
536 <        bare  & UA hexane    & 46.5(3.2) & 49.4(4.5) \\
537 <              & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
538 <              & AA hexane    & 31.0(1.4) & 29.4(1.3) \\
539 <              & UA toluene   & 70.1(1.3) & 65.8(0.5) \\
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{modelTest}
584 >      \label{spceThermal}
585      \end{center}
586    \end{minipage}
587   \end{table*}
588  
589 < To facilitate direct comparison between force fields, systems with the
590 < same capping agent and solvent were prepared with the same length
591 < scales for the simulation cells.
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  
603 < On bare metal / solvent surfaces, different force field models for
604 < hexane yield similar results for both $G$ and $G^\prime$, and these
605 < two definitions agree with each other very well. This is primarily an
606 < indicator of weak interactions between the metal and the solvent.
603 > \begin{table*}
604 >  \begin{minipage}{\linewidth}
605 >    \begin{center}
606 >      
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}{cccc}
612 >        \hline\hline
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 >        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{spceShear}
629 >    \end{center}
630 >  \end{minipage}
631 > \end{table*}
632  
633 < For the fully-covered surfaces, the choice of force field for the
634 < capping agent and solvent has a large impact on the calculated values
635 < of $G$ and $G^\prime$.  The AA thiol to AA solvent conductivities are
636 < much larger than their UA to UA counterparts, and these values exceed
637 < the experimental estimates by a large measure.  The AA force field
638 < allows significant energy to go into C-H (or C-D) stretching modes,
639 < and since these modes are high frequency, this non-quantum behavior is
640 < likely responsible for the overestimate of the conductivity.  Compared
641 < to the AA model, the UA model yields more reasonable conductivity
642 < values with much higher computational efficiency.
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 > 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{Are electronic excitations in the metal important?}
659 < Because they lack electronic excitations, the QSC and related embedded
660 < atom method (EAM) models for gold are known to predict unreasonably
661 < low values for bulk conductivity
662 < ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
663 < conductance between the phases ($G$) is governed primarily by phonon
664 < excitation (and not electronic degrees of freedom), one would expect a
665 < classical model to capture most of the interfacial thermal
666 < conductance.  Our results for $G$ and $G^\prime$ indicate that this is
667 < indeed the case, and suggest that the modeling of interfacial thermal
668 < transport depends primarily on the description of the interactions
669 < between the various components at the interface.  When the metal is
670 < chemically capped, the primary barrier to thermal conductivity appears
671 < to be the interface between the capping agent and the surrounding
672 < solvent, so the excitations in the metal have little impact on the
673 < value of $G$.
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 < \subsection{Effects due to methodology and simulation parameters}
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 < We have varied the parameters of the simulations in order to
716 < investigate how these factors would affect the computation of $G$.  Of
717 < particular interest are: 1) the length scale for the applied thermal
718 < gradient (modified by increasing the amount of solvent in the system),
719 < 2) the sign and magnitude of the applied thermal flux, 3) the average
720 < temperature of the simulation (which alters the solvent density during
721 < equilibration), and 4) the definition of the interfacial conductance
722 < (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
723 < calculation.
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  
596 Systems of different lengths were prepared by altering the number of
597 solvent molecules and extending the length of the box along the $z$
598 axis to accomodate the extra solvent.  Equilibration at the same
599 temperature and pressure conditions led to nearly identical surface
600 areas ($L_x$ and $L_y$) available to the metal and capping agent,
601 while the extra solvent served mainly to lengthen the axis that was
602 used to apply the thermal flux.  For a given value of the applied
603 flux, the different $z$ length scale has only a weak effect on the
604 computed conductivities.
605
606 \subsubsection{Effects of applied flux}
607 The NIVS algorithm allows changes in both the sign and magnitude of
608 the applied flux.  It is possible to reverse the direction of heat
609 flow simply by changing the sign of the flux, and thermal gradients
610 which would be difficult to obtain experimentally ($5$ K/\AA) can be
611 easily simulated.  However, the magnitude of the applied flux is not
612 arbitrary if one aims to obtain a stable and reliable thermal gradient.
613 A temperature gradient can be lost in the noise if $|J_z|$ is too
614 small, and excessive $|J_z|$ values can cause phase transitions if the
615 extremes of the simulation cell become widely separated in
616 temperature.  Also, if $|J_z|$ is too large for the bulk conductivity
617 of the materials, the thermal gradient will never reach a stable
618 state.  
619
620 Within a reasonable range of $J_z$ values, we were able to study how
621 $G$ changes as a function of this flux.  In what follows, we use
622 positive $J_z$ values to denote the case where energy is being
623 transferred by the method from the metal phase and into the liquid.
624 The resulting gradient therefore has a higher temperature in the
625 liquid phase.  Negative flux values reverse this transfer, and result
626 in higher temperature metal phases.  The conductance measured under
627 different applied $J_z$ values is listed in Tables 2 and 3 in the
628 supporting information. These results do not indicate that $G$ depends
629 strongly on $J_z$ within this flux range. The linear response of flux
630 to thermal gradient simplifies our investigations in that we can rely
631 on $G$ measurement with only a small number $J_z$ values.
632
633 The sign of $J_z$ is a different matter, however, as this can alter
634 the temperature on the two sides of the interface. The average
635 temperature values reported are for the entire system, and not for the
636 liquid phase, so at a given $\langle T \rangle$, the system with
637 positive $J_z$ has a warmer liquid phase.  This means that if the
638 liquid carries thermal energy via diffusive transport, {\it positive}
639 $J_z$ values will result in increased molecular motion on the liquid
640 side of the interface, and this will increase the measured
641 conductivity.
642
643 \subsubsection{Effects due to average temperature}
644
645 We also studied the effect of average system temperature on the
646 interfacial conductance.  The simulations are first equilibrated in
647 the NPT ensemble at 1 atm.  The TraPPE-UA model for hexane tends to
648 predict a lower boiling point (and liquid state density) than
649 experiments.  This lower-density liquid phase leads to reduced contact
650 between the hexane and butanethiol, and this accounts for our
651 observation of lower conductance at higher temperatures.  In raising
652 the average temperature from 200K to 250K, the density drop of
653 $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
654 conductance.
655
656 Similar behavior is observed in the TraPPE-UA model for toluene,
657 although this model has better agreement with the experimental
658 densities of toluene.  The expansion of the toluene liquid phase is
659 not as significant as that of the hexane (8.3\% over 100K), and this
660 limits the effect to $\sim$20\% drop in thermal conductivity.
661
662 Although we have not mapped out the behavior at a large number of
663 temperatures, is clear that there will be a strong temperature
664 dependence in the interfacial conductance when the physical properties
665 of one side of the interface (notably the density) change rapidly as a
666 function of temperature.
667
668 Besides the lower interfacial thermal conductance, surfaces at
669 relatively high temperatures are susceptible to reconstructions,
670 particularly when butanethiols fully cover the Au(111) surface. These
671 reconstructions include surface Au atoms which migrate outward to the
672 S atom layer, and butanethiol molecules which embed into the surface
673 Au layer. The driving force for this behavior is the strong Au-S
674 interactions which are modeled here with a deep Lennard-Jones
675 potential. This phenomenon agrees with reconstructions that have been
676 experimentally
677 observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}.  Vlugt
678 {\it et al.} kept their Au(111) slab rigid so that their simulations
679 could reach 300K without surface
680 reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
681 blur the interface, the measurement of $G$ becomes more difficult to
682 conduct at higher temperatures.  For this reason, most of our
683 measurements are undertaken at $\langle T\rangle\sim$200K where
684 reconstruction is minimized.
685
686 However, when the surface is not completely covered by butanethiols,
687 the simulated system appears to be more resistent to the
688 reconstruction. Our Au / butanethiol / toluene system had the Au(111)
689 surfaces 90\% covered by butanethiols, but did not see this above
690 phenomena even at $\langle T\rangle\sim$300K.  That said, we did
691 observe butanethiols migrating to neighboring three-fold sites during
692 a simulation.  Since the interface persisted in these simulations, we
693 were able to obtain $G$'s for these interfaces even at a relatively
694 high temperature without being affected by surface reconstructions.
695
696 \section{Discussion}
697 [COMBINE W. RESULTS]
698 The primary result of this work is that the capping agent acts as an
699 efficient thermal coupler between solid and solvent phases.  One of
700 the ways the capping agent can carry out this role is to down-shift
701 between the phonon vibrations in the solid (which carry the heat from
702 the gold) and the molecular vibrations in the liquid (which carry some
703 of the heat in the solvent).
704
705 To investigate the mechanism of interfacial thermal conductance, the
706 vibrational power spectrum was computed. Power spectra were taken for
707 individual components in different simulations. To obtain these
708 spectra, simulations were run after equilibration in the
709 microcanonical (NVE) ensemble and without a thermal
710 gradient. Snapshots of configurations were collected at a frequency
711 that is higher than that of the fastest vibrations occurring in the
712 simulations. With these configurations, the velocity auto-correlation
713 functions can be computed:
714 \begin{equation}
715 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
716 \label{vCorr}
717 \end{equation}
718 The power spectrum is constructed via a Fourier transform of the
719 symmetrized velocity autocorrelation function,
720 \begin{equation}
721  \hat{f}(\omega) =
722  \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
723 \label{fourier}
724 \end{equation}
725
726 \subsection{The role of specific vibrations}
727 The vibrational spectra for gold slabs in different environments are
728 shown as in Figure \ref{specAu}. Regardless of the presence of
729 solvent, the gold surfaces which are covered by butanethiol molecules
730 exhibit an additional peak observed at a frequency of
731 $\sim$165cm$^{-1}$.  We attribute this peak to the S-Au bonding
732 vibration. This vibration enables efficient thermal coupling of the
733 surface Au layer to the capping agents. Therefore, in our simulations,
734 the Au / S interfaces do not appear to be the primary barrier to
735 thermal transport when compared with the butanethiol / solvent
736 interfaces.  This supports the results of Luo {\it et
737  al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
738 twice as large as what we have computed for the thiol-liquid
739 interfaces.
740
741 \begin{figure}
742 \includegraphics[width=\linewidth]{vibration}
743 \caption{The vibrational power spectrum for thiol-capped gold has an
744  additional vibrational peak at $\sim $165cm$^{-1}$.  Bare gold
745  surfaces (both with and without a solvent over-layer) are missing
746  this peak.   A similar peak at  $\sim $165cm$^{-1}$ also appears in
747  the vibrational power spectrum for the butanethiol capping agents.}
748 \label{specAu}
749 \end{figure}
750
751 Also in this figure, we show the vibrational power spectrum for the
752 bound butanethiol molecules, which also exhibits the same
753 $\sim$165cm$^{-1}$ peak.
754
755 \subsection{Overlap of power spectra}
756 A comparison of the results obtained from the two different organic
757 solvents can also provide useful information of the interfacial
758 thermal transport process.  In particular, the vibrational overlap
759 between the butanethiol and the organic solvents suggests a highly
760 efficient thermal exchange between these components.  Very high
761 thermal conductivity was observed when AA models were used and C-H
762 vibrations were treated classically. The presence of extra degrees of
763 freedom in the AA force field yields higher heat exchange rates
764 between the two phases and results in a much higher conductivity than
765 in the UA force field. The all-atom classical models include high
766 frequency modes which should be unpopulated at our relatively low
767 temperatures.  This artifact is likely the cause of the high thermal
768 conductance in all-atom MD simulations.
769
770 The similarity in the vibrational modes available to solvent and
771 capping agent can be reduced by deuterating one of the two components
772 (Fig. \ref{aahxntln}).  Once either the hexanes or the butanethiols
773 are deuterated, one can observe a significantly lower $G$ and
774 $G^\prime$ values (Table \ref{modelTest}).
775
776 \begin{figure}
777 \includegraphics[width=\linewidth]{aahxntln}
778 \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
779  systems. When butanethiol is deuterated (lower left), its
780  vibrational overlap with hexane decreases significantly.  Since
781  aromatic molecules and the butanethiol are vibrationally dissimilar,
782  the change is not as dramatic when toluene is the solvent (right).}
783 \label{aahxntln}
784 \end{figure}
785
786 For the Au / butanethiol / toluene interfaces, having the AA
787 butanethiol deuterated did not yield a significant change in the
788 measured conductance. Compared to the C-H vibrational overlap between
789 hexane and butanethiol, both of which have alkyl chains, the overlap
790 between toluene and butanethiol is not as significant and thus does
791 not contribute as much to the heat exchange process.
792
793 Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
794 that the {\it intra}molecular heat transport due to alkylthiols is
795 highly efficient.  Combining our observations with those of Zhang {\it
796  et al.}, it appears that butanethiol acts as a channel to expedite
797 heat flow from the gold surface and into the alkyl chain.  The
798 vibrational coupling between the metal and the liquid phase can
799 therefore be enhanced with the presence of suitable capping agents.
800
801 Deuterated models in the UA force field did not decouple the thermal
802 transport as well as in the AA force field.  The UA models, even
803 though they have eliminated the high frequency C-H vibrational
804 overlap, still have significant overlap in the lower-frequency
805 portions of the infrared spectra (Figure \ref{uahxnua}).  Deuterating
806 the UA models did not decouple the low frequency region enough to
807 produce an observable difference for the results of $G$ (Table
808 \ref{modelTest}).
809
810 \begin{figure}
811 \includegraphics[width=\linewidth]{uahxnua}
812 \caption{Vibrational power spectra for UA models for the butanethiol
813  and hexane solvent (upper panel) show the high degree of overlap
814  between these two molecules, particularly at lower frequencies.
815  Deuterating a UA model for the solvent (lower panel) does not
816  decouple the two spectra to the same degree as in the AA force
817  field (see Fig \ref{aahxntln}).}
818 \label{uahxnua}
819 \end{figure}
820
726   \section{Conclusions}
727 < The NIVS algorithm has been applied to simulations of
728 < butanethiol-capped Au(111) surfaces in the presence of organic
729 < solvents. This algorithm allows the application of unphysical thermal
730 < flux to transfer heat between the metal and the liquid phase. With the
731 < 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.
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 < Our simulations have seen significant conductance enhancement in the
737 < presence of capping agent, compared with the bare gold / liquid
738 < interfaces. The vibrational coupling between the metal and the liquid
739 < phase is enhanced by a chemically-bonded capping agent. Furthermore,
740 < the coverage percentage of the capping agent plays an important role
836 < in the interfacial thermal transport process. Moderately low coverages
837 < allow higher contact between capping agent and solvent, and thus could
838 < further enhance the heat transfer process, giving a non-monotonic
839 < behavior of conductance with increasing coverage.
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 results, particularly using the UA models, agree well with
743 < available experimental data.  The AA models tend to overestimate the
744 < interfacial thermal conductance in that the classically treated C-H
745 < vibrations become too easily populated. Compared to the AA models, the
746 < UA models have higher computational efficiency with satisfactory
747 < accuracy, and thus are preferable in modeling interfacial thermal
847 < transport.
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  
849 Of the two definitions for $G$, the discrete form
850 (Eq. \ref{discreteG}) was easier to use and gives out relatively
851 consistent results, while the derivative form (Eq. \ref{derivativeG})
852 is not as versatile. Although $G^\prime$ gives out comparable results
853 and follows similar trend with $G$ when measuring close to fully
854 covered or bare surfaces, the spatial resolution of $T$ profile
855 required for the use of a derivative form is limited by the number of
856 bins and the sampling required to obtain thermal gradient information.
857
858 Vlugt {\it et al.} have investigated the surface thiol structures for
859 nanocrystalline gold and pointed out that they differ from those of
860 the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
861 difference could also cause differences in the interfacial thermal
862 transport behavior. To investigate this problem, one would need an
863 effective method for applying thermal gradients in non-planar
864 (i.e. spherical) geometries.
865
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
# Line 875 | Line 758 | Dame.
758  
759   \end{doublespace}
760   \end{document}
878

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines