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

Comparing trunk/openmdDocs/openmdDoc.tex (file contents):
Revision 3607 by gezelter, Wed Jun 16 15:38:45 2010 UTC vs.
Revision 3793 by skuang, Mon Aug 20 21:03:35 2012 UTC

# Line 17 | Line 17
17   \textwidth 6.5in
18   \brokenpenalty=10000
19   \renewcommand{\baselinestretch}{1.2}
20 + \usepackage[square, comma, sort&compress]{natbib}
21 + \bibpunct{[}{]}{,}{n}{}{;}
22  
23 +
24   %\renewcommand\citemid{\ } % no comma in optional reference note
25   \lstset{language=C,frame=TB,basicstyle=\tiny,basicstyle=\ttfamily, %
26          xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, %
# Line 38 | Line 41
41   \newcolumntype{H}{p{0.75in}}
42   \newcolumntype{I}{p{5in}}
43  
44 + \newcolumntype{J}{p{1.5in}}
45 + \newcolumntype{K}{p{1.2in}}
46 + \newcolumntype{L}{p{1.5in}}
47 + \newcolumntype{M}{p{1.55in}}
48  
42 \title{{\sc OpenMD}: Molecular Dynamics in the Open}
49  
50 < \author{Shenyu Kuang, Chunlei Li, Charles F. Vardeman II, \\
45 < Teng Lin, Christopher J. Fennell,  Xiuquan Sun, \\
46 < Kyle Daily, Yang Zheng, Matthew A. Meineke, and J. Daniel Gezelter\\
47 < Department of Chemistry and Biochemistry\\
48 < University of Notre Dame\\
49 < Notre Dame, Indiana 46556}
50 > \title{{\sc OpenMD-2}: Molecular Dynamics in the Open}
51  
52 + \author{Shenyu Kuang, Joseph Michalka, Kelsey Stocker, James Marr, \\
53 +  Teng Lin, Charles F. Vardeman II, Christopher J. Fennell, Xiuquan Sun, \\
54 +  Chunlei Li, Kyle Daily, Yang Zheng, Matthew A. Meineke, and \\
55 +  J. Daniel Gezelter \\
56 +  Department of Chemistry and Biochemistry\\
57 +  University of Notre Dame\\
58 +  Notre Dame, Indiana 46556}
59 +
60   \maketitle
61  
62   \section*{Preface}
# Line 64 | Line 73 | that is easy to learn.
73   that is easy to learn.
74  
75   \tableofcontents
76 < %\listoffigures
77 < %\listoftables
76 > \listoffigures
77 > \listoftables
78  
79   \mainmatter
80  
# Line 490 | Line 499 | fs}^{-1}$), and body-fixed moments of inertia ($\mbox{
499   \endhead
500   \hline
501   \endfoot
502 < {\tt forceField} & string & Sets the force field. & Possible force
503 < fields are DUFF, WATER, LJ, EAM, SC, and CLAY. \\
502 > {\tt forceField} & string & Sets the base name for the force field file &
503 > OpenMD appends a {\tt .frc} to the end of this to look for a force
504 > field file.\\
505   {\tt component} & & Defines the molecular components of the system &
506   Every {\tt $<$MetaData$>$} block must have a component statement. \\
507   {\tt minimizer} & string & Chooses a minimizer & Possible minimizers
508   are SD and CG. Either {\tt ensemble} or {\tt minimizer} must be specified. \\
509   {\tt ensemble} & string & Sets the ensemble. & Possible ensembles are
510 < NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, and LD.  Either {\tt ensemble}
510 > NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, LD and LangevinHull.  Either {\tt ensemble}
511   or {\tt minimizer} must be specified. \\
512   {\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be
513   small enough to sample the fastest motion of the simulation. ({\tt
# Line 707 | Line 717 | molecule{
717    <MetaData>
718   molecule{
719    name = "I2";
720 <  atom[0]{
721 <    type = "I";
722 <  }
713 <  atom[1]{
714 <    type = "I";
715 <  }
716 <  bond{
717 <    members( 0, 1);
718 <  }
720 >  atom[0]{ type = "I"; }
721 >  atom[1]{ type = "I"; }
722 >  bond{ members( 0, 1); }
723   }
724   molecule{
725    name = "HCl"
726 <  atom[0]{
727 <    type = "H";
728 <  }
725 <  atom[1]{
726 <    type = "Cl";
727 <  }
728 <  bond{
729 <    members( 0, 1);
730 <  }
726 >  atom[0]{ type = "H";}
727 >  atom[1]{ type = "Cl";}
728 >  bond{ members( 0, 1); }
729   }
730   component{
731    type = "HCl";
# Line 1906 | Line 1904 | LD & Langevin Dynamics & {\tt ensemble = LD;} \\
1904   &  (with separate barostats on each box dimension) & \\
1905   LD & Langevin Dynamics & {\tt ensemble = LD;} \\
1906   &  (approximates the effects of an implicit solvent) & \\
1907 + LangevinHull & Non-periodic Langevin Dynamics  & {\tt ensemble = LangevinHull;} \\
1908 + & (Langevin Dynamics for molecules on convex hull;\\
1909 + & Newtonian for interior molecules) & \\
1910   \end{tabular}
1911   \end{center}
1912  
# Line 2391 | Line 2392 | ${\bf V} =
2392   in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
2393   ${\bf V} =
2394   \left\{{\bf v},{\bf \omega}\right\}$. The right side of
2395 < Eq.~\ref{LDGeneralizedForm} consists of three generalized forces: a
2395 > Eq. \ref{LDGeneralizedForm} consists of three generalized forces: a
2396   system force (${\bf F}_{s}$), a frictional or dissipative force (${\bf
2397   F}_{f}$) and a stochastic force (${\bf F}_{r}$). While the evolution
2398   of the system in Newtonian mechanics is typically done in the lab
# Line 2613 | Line 2614 | program that is included in the {\sc OpenMD} distribut
2614   \endhead
2615   \hline
2616   \endfoot
2617 < {\tt viscosity} & centipoise & Sets the value of viscosity of the implicit
2617 > {\tt viscosity} & poise & Sets the value of viscosity of the implicit
2618   solvent  \\
2619   {\tt targetTemp} & K & Sets the target temperature of the system.
2620   This parameter must be specified to use Langevin dynamics. \\
2621   {\tt HydroPropFile} & string & Specifies the name of the resistance
2622   tensor (usually a {\tt .diff} file) which is precalculated by {\tt
2623 < Hydro}. This keyworkd is not necessary if the simulation contains only
2623 > Hydro}. This keyword is not necessary if the simulation contains only
2624   simple bodies (spheres and ellipsoids). \\
2625   {\tt beadSize} & $\mbox{\AA}$ & Sets the diameter of the beads to use
2626   when the {\tt RoughShell} model is used to approximate the resistance
2627   tensor.
2628   \label{table:ldParameters}
2629 + \end{longtable}
2630 +
2631 + \section{Constant Pressure without periodic boundary conditions (The LangevinHull)}
2632 +
2633 + The Langevin Hull\cite{Vardeman2011} uses an external bath at a fixed constant pressure
2634 + ($P$) and temperature ($T$) with an effective solvent viscosity
2635 + ($\eta$).  This bath interacts only with the objects on the exterior
2636 + hull of the system.  Defining the hull of the atoms in a simulation is
2637 + done in a manner similar to the approach of Kohanoff, Caro and
2638 + Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous configuration
2639 + of the atoms in the system is considered as a point cloud in three
2640 + dimensional space.  Delaunay triangulation is used to find all facets
2641 + between coplanar
2642 + neighbors.\cite{delaunay,springerlink:10.1007/BF00977785} In highly
2643 + symmetric point clouds, facets can contain many atoms, but in all but
2644 + the most symmetric of cases, the facets are simple triangles in
2645 + 3-space which contain exactly three atoms.
2646 +
2647 + The convex hull is the set of facets that have {\it no concave
2648 +  corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
2649 + eliminates all facets on the interior of the point cloud, leaving only
2650 + those exposed to the bath. Sites on the convex hull are dynamic; as
2651 + molecules re-enter the cluster, all interactions between atoms on that
2652 + molecule and the external bath are removed.  Since the edge is
2653 + determined dynamically as the simulation progresses, no {\it a priori}
2654 + geometry is defined. The pressure and temperature bath interacts only
2655 + with the atoms on the edge and not with atoms interior to the
2656 + simulation.
2657 +
2658 + Atomic sites in the interior of the simulation move under standard
2659 + Newtonian dynamics,
2660 + \begin{equation}
2661 + m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
2662 + \label{eq:Newton}
2663 + \end{equation}
2664 + where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
2665 + instantaneous velocity of site $i$ at time $t$, and $U$ is the total
2666 + potential energy.  For atoms on the exterior of the cluster
2667 + (i.e. those that occupy one of the vertices of the convex hull), the
2668 + equation of motion is modified with an external force, ${\mathbf
2669 +  F}_i^{\mathrm ext}$:
2670 + \begin{equation}
2671 + m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
2672 + \end{equation}
2673 +
2674 + The external bath interacts indirectly with the atomic sites through
2675 + the intermediary of the hull facets.  Since each vertex (or atom)
2676 + provides one corner of a triangular facet, the force on the facets are
2677 + divided equally to each vertex.  However, each vertex can participate
2678 + in multiple facets, so the resultant force is a sum over all facets
2679 + $f$ containing vertex $i$:
2680 + \begin{equation}
2681 + {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
2682 +    } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\  {\mathbf
2683 +  F}_f^{\mathrm ext}
2684 + \end{equation}
2685 +
2686 + The external pressure bath applies a force to the facets of the convex
2687 + hull in direct proportion to the area of the facet, while the thermal
2688 + coupling depends on the solvent temperature, viscosity and the size
2689 + and shape of each facet. The thermal interactions are expressed as a
2690 + standard Langevin description of the forces,
2691 + \begin{equation}
2692 + \begin{array}{rclclcl}
2693 + {\mathbf F}_f^{\text{ext}} & = &  \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
2694 + & = &  -\hat{n}_f P A_f  & - & \Xi_f(t) {\mathbf v}_f(t)  & + & {\mathbf R}_f(t)
2695 + \end{array}
2696 + \end{equation}
2697 + Here, $A_f$ and $\hat{n}_f$ are the area and (outward-facing) normal
2698 + vectors for facet $f$, respectively.  ${\mathbf v}_f(t)$ is the
2699 + velocity of the facet centroid,
2700 + \begin{equation}
2701 + {\mathbf v}_f(t) =  \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
2702 + \end{equation}
2703 + and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
2704 + depends on the geometry and surface area of facet $f$ and the
2705 + viscosity of the bath.  The resistance tensor is related to the
2706 + fluctuations of the random force, $\mathbf{R}(t)$, by the
2707 + fluctuation-dissipation theorem (see Eq. \ref{eq:randomForce}).
2708 +
2709 + Once the resistance tensor is known for a given facet, a stochastic
2710 + vector that has the properties in Eq. (\ref{eq:randomForce}) can be
2711 + calculated efficiently by carrying out a Cholesky decomposition to
2712 + obtain the square root matrix of the resistance tensor (see
2713 + Eq. \ref{eq:Cholesky}).
2714 +
2715 + Our treatment of the resistance tensor for the Langevin Hull facets is
2716 + approximate.  $\Xi_f$ for a rigid triangular plate would normally be
2717 + treated as a $6 \times 6$ tensor that includes translational and
2718 + rotational drag as well as translational-rotational coupling. The
2719 + computation of resistance tensors for rigid bodies has been detailed
2720 + elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
2721 + but the standard approach involving bead approximations would be
2722 + prohibitively expensive if it were recomputed at each step in a
2723 + molecular dynamics simulation.
2724 +
2725 + Instead, we are utilizing an approximate resistance tensor obtained by
2726 + first constructing the Oseen tensor for the interaction of the
2727 + centroid of the facet ($f$) with each of the subfacets $\ell=1,2,3$,
2728 + \begin{equation}
2729 + T_{\ell f}=\frac{A_\ell}{8\pi\eta R_{\ell f}}\left(I +
2730 +  \frac{\mathbf{R}_{\ell f}\mathbf{R}_{\ell f}^T}{R_{\ell f}^2}\right)
2731 + \end{equation}
2732 + Here, $A_\ell$ is the area of subfacet $\ell$ which is a triangle
2733 + containing two of the vertices of the facet along with the centroid.
2734 + $\mathbf{R}_{\ell f}$ is the vector between the centroid of facet $f$
2735 + and the centroid of sub-facet $\ell$, and $I$ is the ($3 \times 3$)
2736 + identity matrix.  $\eta$ is the viscosity of the external bath.
2737 +
2738 + The tensors for each of the sub-facets are added together, and the
2739 + resulting matrix is inverted to give a $3 \times 3$ resistance tensor
2740 + for translations of the triangular facet,
2741 + \begin{equation}
2742 + \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
2743 + \end{equation}
2744 + Note that this treatment ignores rotations (and
2745 + translational-rotational coupling) of the facet.  In compact systems,
2746 + the facets stay relatively fixed in orientation between
2747 + configurations, so this appears to be a reasonably good approximation.
2748 +
2749 + At each
2750 + molecular dynamics time step, the following process is carried out:
2751 + \begin{enumerate}
2752 + \item The standard inter-atomic forces ($\nabla_iU$) are computed.
2753 + \item Delaunay triangulation is carried out using the current atomic
2754 +  configuration.
2755 + \item The convex hull is computed and facets are identified.
2756 + \item For each facet:
2757 + \begin{itemize}
2758 + \item[a.] The force from the pressure bath ($-\hat{n}_fPA_f$) is
2759 +  computed.
2760 + \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
2761 +  viscosity ($\eta$) of the bath.
2762 + \item[c.] Facet drag ($-\Xi_f(t) \mathbf{v}_f(t)$) forces are
2763 +  computed.
2764 + \item[d.] Random forces ($\mathbf{R}_f(t)$) are computed using the
2765 +  resistance tensor and the temperature ($T$) of the bath.
2766 + \end{itemize}
2767 + \item The facet forces are divided equally among the vertex atoms.
2768 + \item Atomic positions and velocities are propagated.
2769 + \end{enumerate}
2770 + The Delaunay triangulation and computation of the convex hull are done
2771 + using calls to the qhull library,\cite{Qhull} and for this reason, if
2772 + qhull is not detected during the build, the Langevin Hull integrator
2773 + will not be available.  There is a minimal penalty for computing the
2774 + convex hull and resistance tensors at each step in the molecular
2775 + dynamics simulation (roughly 0.02 $\times$ cost of a single force
2776 + evaluation).
2777 +
2778 + \begin{longtable}[c]{GBF}
2779 + \caption{Meta-data Keywords: Required parameters for the Langevin Hull integrator}
2780 + \\
2781 + {\bf keyword} & {\bf units} & {\bf use}  \\ \hline
2782 + \endhead
2783 + \hline
2784 + \endfoot
2785 + {\tt viscosity} & poise & Sets the value of viscosity of the implicit
2786 + solven . \\
2787 + {\tt targetTemp} & K & Sets the target temperature of the system.
2788 + This parameter must be specified to use Langevin Hull dynamics. \\
2789 + {\tt targetPressure} & atm & Sets the target pressure of the system.
2790 + This parameter must be specified to use Langevin Hull dynamics. \\
2791 + {\tt usePeriodicBoundaryConditions} & logical & Turns off periodic boundary conditions.
2792 + This parameter must be set to \tt false \\
2793 + \label{table:lhullParameters}
2794   \end{longtable}
2795  
2796 +
2797   \section{\label{sec:constraints}Constraint Methods}
2798  
2799   \subsection{\label{section:rattle}The {\sc rattle} Method for Bond
# Line 2775 | Line 2942 | Harmonic Forces are used by default
2942   \label{table:zconParams}
2943   \end{longtable}
2944  
2945 < \chapter{\label{section:restraints}Restraints}
2946 < Restraints are external potentials that are added to a system to keep
2947 < particular molecules or collections of particles close to some
2948 < reference structure.  A restraint can be a collective
2945 > % \chapter{\label{section:restraints}Restraints}
2946 > % Restraints are external potentials that are added to a system to
2947 > % keep particular molecules or collections of particles close to some
2948 > % reference structure.  A restraint can be a collective
2949  
2950   \chapter{\label{section:thermInt}Thermodynamic Integration}
2951  
# Line 2916 | Line 3083 | Einstein crystal
3083   \mbox{rad}^{-2}$ & & spring constant for rotation around z-axis in
3084   Einstein crystal
3085   \label{table:thermIntParams}
3086 + \end{longtable}
3087 +
3088 + \chapter{\label{section:rnemd}RNEMD}
3089 +
3090 + There are many ways to compute transport properties from molecular
3091 + dynamics simulations.  Equilibrium Molecular Dynamics (EMD)
3092 + simulations can be used by computing relevant time correlation
3093 + functions and assuming linear response theory holds.  These approaches
3094 + are generally subject to noise and poor convergence of the relevant
3095 + correlation functions. Traditional Non-equilibrium Molecular Dynamics
3096 + (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
3097 + simulation.  However, the resulting flux is often difficult to
3098 + measure. Furthermore, problems arise for NEMD simulations of
3099 + heterogeneous systems, such as phase-phase boundaries or interfaces,
3100 + where the type of gradient to enforce at the boundary between
3101 + materials is unclear.
3102 +
3103 + {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
3104 + a different approach in that an unphysical {\it flux} is imposed
3105 + between different regions or ``slabs'' of the simulation box.  The
3106 + response of the system is to develop a temperature or momentum {\it
3107 +  gradient} between the two regions. Since the amount of the applied
3108 + flux is known exactly, and the measurement of gradient is generally
3109 + less complicated, imposed-flux methods typically take shorter
3110 + simulation times to obtain converged results for transport properties.
3111 +
3112 + \begin{figure}
3113 + \includegraphics[width=\linewidth]{rnemdDemo}
3114 + \caption{The (VSS) RNEMD approach imposes unphysical transfer of both
3115 +  linear momentum and kinetic energy between a ``hot'' slab and a
3116 +  ``cold'' slab in the simulation box.  The system responds to this
3117 +  imposed flux by generating both momentum and temperature gradients.
3118 +  The slope of the gradients can then be used to compute transport
3119 +  properties (e.g. shear viscosity and thermal conductivity).}
3120 + \label{rnemdDemo}
3121 + \end{figure}
3122 +
3123 + The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
3124 +  al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
3125 + as a sequence of imaginary elastic collisions between particles in
3126 + opposite slabs.  In each collision, the entire momentum vectors of
3127 + both particles may be exchanged to generate a thermal
3128 + flux. Alternatively, a single component of the momentum vectors may be
3129 + exchanged to generate a shear flux.  This algorithm turns out to be
3130 + quite useful in many simulations. However, the M\"{u}ller-Plathe
3131 + swapping approach perturbs the system away from ideal
3132 + Maxwell-Boltzmann distributions, and this may leads to undesirable
3133 + side-effects when the applied flux becomes large.\cite{Maginn:2010}
3134 + This limits the application of the swapping algorithm, so in OpenMD,
3135 + we implement two additional algorithms for RNEMD in addition to the
3136 + original swapping approach.
3137 +
3138 + {\bf Non-Isotropic Velocity Scaling (NIVS):}\cite{kuang:164101}
3139 + Instead of having momentum exchange between {\it individual particles}
3140 + in each slab, the NIVS algorithm applies velocity scaling to all of
3141 + the selected particles in both slabs.  A combination of linear
3142 + momentum, kinetic energy, and flux constraint equations governs the
3143 + amount of velocity scaling performed at each step.  Interested readers
3144 + should consult ref. \citealp{kuang:164101} for further details on the
3145 + methodology.
3146 +
3147 + NIVS has been shown to be very effective at producing thermal
3148 + gradients and for computing thermal conductivities, particularly for
3149 + heterogeneous interfaces.  Although the NIVS algorithm can also be
3150 + applied to impose a directional momentum flux, thermal anisotropy was
3151 + observed in relatively high flux simulations, and the method is not
3152 + suitable for imposing a shear flux.
3153 +
3154 + {\bf Velocity Shearing and Scaling (VSS)}:\cite{2012MolPh.110..691K}
3155 + The third RNEMD algorithm implemented in OpenMD utilizes a series of
3156 + simultaneous velocity shearing and scaling exchanges between the two
3157 + slabs.  This method results in a set of simpler equations to satisfy
3158 + the conservation constraints while creating a desired flux between the
3159 + two slabs.
3160 +
3161 + The VSS approach is versatile in that it may be used to implement both
3162 + thermal and shear transport either separately or simultaneously.
3163 + Perturbations of velocities away from the ideal Maxwell-Boltzmann
3164 + distributions are minimal, and thermal anisotropy is kept to a
3165 + minimum.  This ability to generate simultaneous thermal and shear
3166 + fluxes has been utilized to map out the shear viscosity of SPC/E water
3167 + in a wide range of temperature (90~K) just with a single simulation.
3168 + VSS-RNEMD also allows the directional momentum flux to have quite
3169 + arbitary directions, which could aid in the study of anisotropic solid
3170 + surfaces in contact with liquid environments.
3171 +
3172 + {\bf What the user needs to specify:} To carry out a RNEMD simulation,
3173 + a user must specify a number of parameters in the MetaData (.md) file.
3174 + Because the RNEMD methods have a large number of parameters, these
3175 + must be enclosed in a {\tt RNEMD\{...\}} block.  The most important
3176 + parameters to specify are the {\tt useRNEMD}, {\tt fluxType} and flux
3177 + parameters. Most other parameters (summarized in table
3178 + \ref{table:rnemd}) have reasonable default values.  {\tt fluxType}
3179 + sets up the kind of exchange that will be carried out between the two
3180 + slabs (either Kinetic Energy ({\tt KE}) or momentum ({\tt Px, Py, Pz,
3181 +  Pvector}), or some combination of these).  The flux is specified
3182 + with the use of three possible parameters: {\tt kineticFlux} for
3183 + kinetic energy exchange, as well as {\tt momentumFlux} or {\tt
3184 +  momentumFluxVector} for simulations with directional exchange.
3185 +
3186 + {\bf How to process the results:} OpenMD will generate a {\tt .rnemd}
3187 + file with the same prefix as the original {\tt .md} file.  This file
3188 + contains a running average of properties of interest computed within a
3189 + set of bins that divide the simulation cell along the $z$-axis.  The
3190 + first column of the {\tt .rnemd} file is the $z$ coordinate of the
3191 + center of each bin, while following columns may contain the average
3192 + temperature, velocity, or density within each bin.  The output format
3193 + in the {\tt .rnemd} file can be altered with the {\tt outputFields},
3194 + {\tt outputBins}, and {\tt outputFileName} parameters.  A report at
3195 + the top of the {\tt .rnemd} file contains the current exchange totals
3196 + as well as the average flux applied during the simulation.  Using the
3197 + slope of the temperature or velocity gradient obtaine from the {\tt
3198 +  .rnemd} file along with the applied flux, the user can very simply
3199 + arrive at estimates of thermal conductivities ($\lambda$),
3200 + \begin{equation}
3201 + J_z = -\lambda \frac{\partial T}{\partial z},
3202 + \end{equation}
3203 + and shear viscosities ($\eta$),
3204 + \begin{equation}
3205 + j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
3206 + \end{equation}
3207 + Here, the quantities on the left hand side are the actual flux values
3208 + (in the header of the {\tt .rnemd} file), while the slopes are
3209 + obtained from linear fits to the gradients observed in the {\tt
3210 +  .rnemd} file.
3211 +
3212 + More complicated simulations (including interfaces) require a bit more
3213 + care.  Here the second derivative may be required to compute the
3214 + interfacial thermal conductance,
3215 + \begin{align}
3216 +  G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
3217 +  &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
3218 +      \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
3219 +  &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
3220 +  \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
3221 +  \label{derivativeG}
3222 + \end{align}
3223 + where $z_0$ is the location of the interface between two materials and
3224 + $\mathbf{\hat{n}}$ is a unit vector normal to the interface.  We
3225 + suggest that users interested in interfacial conductance consult
3226 + reference \citealp{kuang:AuThl} for other approaches to computing $G$.
3227 + Users interested in {\it friction coefficients} at heterogeneous
3228 + interfaces may also find reference \citealp{2012MolPh.110..691K}
3229 + useful.
3230 +
3231 + \newpage
3232 +
3233 + \begin{longtable}[c]{JKLM}
3234 + \caption{The following keywords must be enclosed inside a {\tt RNEMD\{\}} block}
3235 + \\
3236 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks}  \\ \hline
3237 + \endhead
3238 + \hline
3239 + \endfoot
3240 + {\tt useRNEMD} & logical & perform RNEMD? & default is ``false'' \\
3241 + {\tt objectSelection} & string & see section \ref{section:syntax}
3242 + for selection syntax & default is ``select all'' \\
3243 + {\tt method} & string & exchange method & one of the following:
3244 + {\tt Swap, NIVS,} or {\tt VSS}  (default is {\tt VSS}) \\
3245 + {\tt fluxType} & string & what is being exchanged between slabs? & one
3246 + of the following: {\tt KE, Px, Py, Pz, Pvector, KE+Px, KE+Py, KE+Pvector} \\
3247 + {\tt kineticFlux} & kcal mol$^{-1}$ \AA$^{-2}$ fs$^{-1}$ & specify the kinetic energy flux &  \\
3248 + {\tt momentumFlux} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux & \\
3249 + {\tt momentumFluxVector} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux when
3250 + {\tt Pvector} is part of the exchange & Vector3d input\\
3251 + {\tt exchangeTime} & fs & how often to perform the exchange & default is 100 fs\\
3252 +
3253 + {\tt slabWidth} & $\mbox{\AA}$ & width of the two exchange slabs & default is $\mathsf{H}_{zz} / 10.0$ \\
3254 + {\tt slabAcenter} & $\mbox{\AA}$ & center of the end slab & default is 0 \\
3255 + {\tt slabBcenter} & $\mbox{\AA}$ & center of the middle slab & default is $\mathsf{H}_{zz} / 2$ \\
3256 + {\tt outputFileName} & string & file name for output histograms & default is the same prefix as the
3257 + .md file, but with the {\tt .rnemd} extension \\
3258 + {\tt outputBins} & int & number of $z$-bins in the output histogram &
3259 + default is 20 \\
3260 + {\tt outputFields} & string & columns to print in the {\tt .rnemd}
3261 + file where each column is separated by a pipe ($\mid$) symbol. & Allowed column names are: {\sc z, temperature, velocity, density} \\
3262 + \label{table:rnemd}
3263   \end{longtable}
3264  
3265  
# Line 3190 | Line 3534 | For example, the phrase {\tt select mass > 16.0 and ch
3534   \end{center}
3535  
3536   For example, the phrase {\tt select mass > 16.0 and charge < -2}
3537 < wouldselect StuntDoubles which have mass greater than 16.0 and charges
3537 > would select StuntDoubles which have mass greater than 16.0 and charges
3538   less than -2.
3539  
3540   \subsection{\label{section:within}Within expressions}
# Line 3592 | Line 3936 | DMR-0079647.
3936   DMR-0079647.
3937  
3938  
3939 < \bibliographystyle{jcc}
3939 > \bibliographystyle{aip}
3940   \bibliography{openmdDoc}
3941  
3942   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines