--- stokes/stokes.tex 2011/12/13 23:31:35 3785 +++ stokes/stokes.tex 2011/12/14 21:59:12 3786 @@ -28,8 +28,8 @@ \begin{document} -\title{A minimal perturbation approach to RNEMD able to simultaneously -impose thermal and momentum gradients} +\title{Velocity Shearing and Scaling RNEMD: a minimally perturbing + method for producing temperature and momentum gradients} \author{Shenyu Kuang and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\ @@ -46,18 +46,19 @@ Notre Dame, Indiana 46556} \begin{abstract} We present a new method for introducing stable nonequilibrium velocity and temperature gradients in molecular dynamics simulations - of heterogeneous systems. This method conserves the linear momentum - and total energy of the system and improves previous Reverse - Non-Equilibrium Molecular Dynamics (RNEMD) methods while maintaining - thermal velocity distributions. It also avoid thermal anisotropy - occured in previous NIVS simulations by using isotropic velocity - scaling on the molecules in specific regions of a system. To test - the method, we have computed the thermal conductivity and shear + of heterogeneous systems. This method conserves both the linear + momentum and total energy of the system and improves previous + reverse non-equilibrium molecular dynamics (RNEMD) methods while + retaining equilibrium thermal velocity distributions in each region + of the system. The new method avoids the thermal anisotropy + produced by previous methods by using isotropic velocity scaling and + shearing on all of the molecules in specific regions. To test the + method, we have computed the thermal conductivity and shear viscosity of model liquid systems as well as the interfacial - frictions of a series of metal/liquid interfaces. Its ability to - combine the thermal and momentum gradients allows us to obtain shear - viscosity data for a range of temperatures in only one trajectory. - + friction coeefficients of a series of solid / liquid interfaces. The + method's ability to combine the thermal and momentum gradients + allows us to obtain shear viscosity data for a range of temperatures + from a single trajectory. \end{abstract} \newpage @@ -69,96 +70,129 @@ Imposed-flux methods in Molecular Dynamics (MD) %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Introduction} -Imposed-flux methods in Molecular Dynamics (MD) -simulations\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101} -can establish steady state systems with an applied flux set vs a -corresponding gradient that can be measured. These methods does not -need many trajectories to provide information of transport properties -of a given system. Thus, they are utilized in computing thermal and -mechanical transfer of homogeneous bulk systems as well as -heterogeneous systems such as solid-liquid +One of the standard ways to compute transport coefficients such as the +viscosities and thermal conductivities of liquids is to use +imposed-flux non-equilibrium molecular dynamics +methods.\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101} +These methods establish stable non-equilibrium conditions in a +simulation box using an applied momentum or thermal flux between +different regions of the box. The corresponding temperature or +velocity gradients which develop in response to the applied flux is +related (via linear response theory) to the transport coefficient of +interest. These methods are quite efficient, in that they do not need +many trajectories to provide information about transport properties. To +date, they have been utilized in computing thermal and mechanical +transfer of both homogeneous liquids as well as heterogeneous systems +such as solid-liquid interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl} -The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that -satisfy linear momentum and total energy conservation of a system when -imposing fluxes in a simulation. Thus they are compatible with various -ensembles, including the micro-canonical (NVE) ensemble, without the -need of an external thermostat. The original approaches proposed by -M\"{u}ller-Plathe {\it et +The reverse non-equilibrium molecular dynamics (RNEMD) methods utilize +additional constraints that ensure conservation of linear momentum and +total energy of the system while imposing the desired flux. The RNEMD +methods are therefore capable of sampling various thermodynamically +relevent ensembles, including the micro-canonical (NVE) ensemble, +without resorting to an external thermostat. The original approaches +proposed by M\"{u}ller-Plathe {\it et al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple -momentum swapping for generating energy/momentum fluxes, which can -also be compatible with particles of different identities. Although -simple to implement in a simulation, this approach can create -nonthermal velocity distributions, as discovered by Tenney and -Maginn\cite{Maginn:2010}. Furthermore, this approach is less efficient -for kinetic energy transfer between particles of different identities, -especially when the mass difference between the particles becomes -significant. This also limits its applications on heterogeneous -interfacial systems. +momentum swapping moves for generating energy/momentum fluxes. The +swapping moves can also be made compatible with particles of different +identities. Although the swapping moves are simple to implement in +molecular simulations, Tenney and Maginn have shown that they create +nonthermal velocity distributions.\cite{Maginn:2010} Furthermore, this +approach is not particularly efficient for kinetic energy transfer +between particles of different identities, particularly when the mass +difference between the particles becomes significant. This problem +makes applying swapping-move RNEMD methods on heterogeneous +interfacial systems somewhat difficult. -Recently, we developed a different approach, using Non-Isotropic -Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose -fluxes. Compared to the momentum swapping move, it scales the velocity -vectors in two separate regions of a simulated system with respective -diagonal scaling matrices. These matrices are determined by solving a -set of equations including linear momentum and kinetic energy -conservation constraints and target flux satisfaction. This method is -able to effectively impose a wide range of kinetic energy fluxes -without obvious perturbation to the velocity distributions of the -simulated systems, regardless of the presence of heterogeneous -interfaces. We have successfully applied this approach in studying the -interfacial thermal conductance at metal-solvent +Recently, we developed a somewhat different approach to applying +thermal fluxes in RNEMD simulation using a Non-Isotropic Velocity +Scaling (NIVS) algorithm.\cite{kuang:164101} This algorithm scales all +atomic velocity vectors in two separate regions of a simulated system +using two diagonal scaling matrices. The scaling matrices are +determined by solving single quartic equation which includes linear +momentum and kinetic energy conservation constraints as well as the +target thermal flux between the regions. The NIVS method is able to +effectively impose a wide range of kinetic energy fluxes without +significant perturbation to the velocity distributions away from ideal +Maxwell-Boltzmann distributions, even in the presence of heterogeneous +interfaces. We successfully applied this approach in studying the +interfacial thermal conductance at chemically-capped metal-solvent interfaces.\cite{kuang:AuThl} -However, the NIVS approach has limited applications in imposing -momentum fluxes. Temperature anisotropy could happen under high -momentum fluxes due to the implementation of this algorithm. Thus, -combining thermal and momentum flux is also difficult to obtain with -this approach. However, such combination may provide a means to -simulate thermal/momentum gradient coupled processes such as Soret -effect in liquid flows. Therefore, developing improved approaches to -extend the applications of the imposed-flux method is desirable. +The NIVS approach works very well for preparing stable {\it thermal} +gradients. However, as we pointed out in the original paper, it has +limited application in imposing {\it linear} momentum fluxes (which +are required for measuring shear viscosities). The reason for this is +that linear momentum flux was being imposed by scaling random +fluctuations of the center of the velocity distributions. Repeated +application of the original NIVS approach resulted in temperature +anisotropy, i.e. the width of the velocity distributions depended on +coordinates perpendicular to the desired gradient direction. For this +reason, combining thermal and momentum fluxes was difficult with the +original NIVS algorithm. However, combinations of thermal and +velocity gradients would provide a means to simulate thermal-linear +coupled processes such as Soret effect in liquid flows. Therefore, +developing improved approaches to the scaling imposed-flux methods +would be useful. -In this paper, we improve the RNEMD methods by proposing a novel -approach to impose fluxes. This approach separate the means of applying -momentum and thermal flux with operations in one time step and thus is -able to simutaneously impose thermal and momentum flux. Furthermore, -the approach retains desirable features of previous RNEMD approaches -and is simpler to implement compared to the NIVS method. In what -follows, we first present the method and its implementation in a -simulation. Then we compare the method on bulk fluids to previous -methods. Also, interfacial frictions are computed for a series of -interfaces. +In this paper, we improve the RNEMD methods by introducing a novel +approach to creating imposed fluxes. This approach separates the +shearing and scaling of the velocity distributions in different +spatial regions and can apply both transformations within a single +time step. The approach retains desirable features of previous RNEMD +approaches and is simpler to implement compared to the earlier NIVS +method. In what follows, we first present the Shearing-and-Scaling +(SS) RNEMD method and its implementation in a simulation. Then we +compare the SS-RNEMD method in bulk fluids to previous methods. We +also compute interfacial frictions are computed for a series of +heterogeneous interfaces. \section{Methodology} -Similar to the NIVS method,\cite{kuang:164101} we consider a -periodic system divided into a series of slabs along a certain axis -(e.g. $z$). The unphysical thermal and/or momentum flux is designated -from the center slab to one of the end slabs, and thus the thermal -flux results in a lower temperature of the center slab than the end -slab, and the momentum flux results in negative center slab momentum -with positive end slab momentum (unless these fluxes are set -negative). Therefore, the center slab is denoted as ``$c$'', while the -end slab as ``$h$''. +In an approach similar to the earlier NIVS method,\cite{kuang:164101} +we consider a periodic system which has been divided into a series of +slabs along a single axis (e.g. $z$). The unphysical thermal and +momentum fluxes are applied from one of the end slabs to the center +slab, and thus the thermal flux produces a higher temperature in the +center slab than in the end slab, and the momentum flux results in a +center slab moving along the positive $y$ axis (see Fig. \ref{cartoon}). +The applied fluxes can be set to negative values if the reverse +gradients are desired. For convenience the center slab is denoted as +the {\it hot} or {\it ``H''} slab, while the end slab is denoted {\it + ``C''} (or {\it cold}). -To impose these fluxes, we periodically apply different set of -operations on velocities of particles {$i$} within the center slab and -those of particles {$j$} within the end slab: +\begin{figure} +\includegraphics[width=\linewidth]{cartoon} +\caption{The SS-RNEMD approach we are introducing imposes unphysical + transfer of both momentum and kinetic energy between a ``hot'' slab + and a ``cold'' slab in the simulation box. The molecular system + responds to this imposed flux by generating both momentum and + temperature gradients. The slope of the gradients can then be used + to compute transport properties (e.g. shear viscosity and thermal + conductivity).} +\label{cartoon} +\end{figure} + +To impose these fluxes, we periodically apply a set of operations on +the velocities of particles {$i$} within the cold slab and a separate +operation on particles {$j$} within the hot slab. \begin{eqnarray} \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\ \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right) \end{eqnarray} -where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes -the instantaneous bulk velocity of slabs $c$ and $h$ respectively -before an operation is applied. When a momentum flux $\vec{j}_z(\vec{p})$ -presents, these bulk velocities would have a corresponding change -($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's -second law: +where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denote +the instantaneous average velocity of the molecules within slabs $C$ +and $H$ respectively. When a momentum flux $\vec{j}_z(\vec{p})$ is +present, these slab-averaged velocities also get corresponding +incremental changes ($\vec{a}_c$ and $\vec{a}_h$ respectively) that +are applied to all particles within each slab. The incremental +changes are obtained using Newton's second law: \begin{eqnarray} -M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\ +M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \label{eq:newton1} \\ M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t +\label{eq:newton2} \end{eqnarray} where $M$ denotes total mass of particles within a slab: \begin{eqnarray} @@ -167,138 +201,142 @@ The above operations already conserve the linear momen \end{eqnarray} and $\Delta t$ is the interval between two separate operations. -The above operations already conserve the linear momentum of a -periodic system. To further satisfy total energy conservation as well -as to impose the thermal flux $J_z$, the following equations are -included as well: -[MAY PUT EXTRA MATH IN SUPPORT INFO OR APPENDIX] +The operations in Eqs. \ref{eq:newton1} and \ref{eq:newton2} already +conserve the linear momentum of a periodic system. To further satisfy +total energy conservation as well as to impose the thermal flux $J_z$, +the following constraint equations must be solved for the two scaling +variables $c$ and $h$: \begin{eqnarray} K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\ K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h -\rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2 +\rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2. +\label{constraint} \end{eqnarray} -where $K_c$ and $K_h$ denotes translational kinetic energy of slabs -$c$ and $h$ respectively before an operation is applied. These -translational kinetic energy conservation equations are sufficient to -ensure total energy conservation, as the operations applied in our -method do not change the kinetic energy related to other degrees of -freedom or the potential energy of a system, given that its potential -energy does not depend on particle velocity. +Here $K_c$ and $K_h$ denote the translational kinetic energy of slabs +$C$ and $H$ respectively. These conservation equations are sufficient +to ensure total energy conservation, as the operations applied in our +method do not change the kinetic energy related to orientational +degrees of freedom or the potential energy of a system (as long as the +potential energy is independent of particle velocity). -The above sets of equations are sufficient to determine the velocity -scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and -$\vec{a}_h$. Note that there are two roots respectively for $c$ and -$h$. However, the positive roots (which are closer to 1) are chosen so -that the perturbations to a system can be reduced to a minimum. Figure -\ref{method} illustrates the implementation sketch of this algorithm -in an individual step. +Equations \ref{eq:newton1}-\ref{constraint} are sufficient to +determine the velocity scaling coefficients ($c$ and $h$) as well as +$\vec{a}_c$ and $\vec{a}_h$. Note that there are usually two roots +respectively for $c$ and $h$. However, the positive roots (which are +closer to 1) are chosen so that the perturbations to the system are +minimal. Figure \ref{method} illustrates the implementation of this +algorithm in an individual step. \begin{figure} -\includegraphics[width=\linewidth]{method} -\caption{Illustration of the implementation of the algorithm in a - single step. Starting from an ideal velocity distribution, the - transformation is used to apply the effect of both a thermal and a - momentum flux from the ``c'' slab to the ``h'' slab. As the figure - shows, thermal distributions can preserve after this operation.} +\centering +\includegraphics[width=5in]{method} +\caption{Illustration of a single step implementation of the + algorithm. Starting with the velocity distributions for the two + slabs in a shearing fluid, the transformation is used to apply the + effect of both a thermal and a momentum flux from the ``c'' slab to + the ``h'' slab. As the figure shows, gaussian distributions are + preserved by both the scaling and shearing operations.} \label{method} \end{figure} -By implementing these operations at a certain frequency, a steady -thermal and/or momentum flux can be applied and the corresponding -temperature and/or momentum gradients can be established. +By implementing these operations at a fixed frequency, stable thermal +and momentum fluxes can both be applied and the corresponding +temperature and momentum gradients can be established. -Compared to the previous NIVS method, this approach is computationally -more efficient in that only quadratic equations are involved to -determine a set of scaling coefficients, while the NIVS method needs -to solve quartic equations. Furthermore, this method implements -isotropic scaling of velocities in respective slabs, unlike the NIVS, -where an extra criteria function is necessary to choose a set of -coefficients that performs a scaling as isotropic as possible. More -importantly, separating the means of momentum flux imposing from -velocity scaling avoids the underlying cause to thermal anisotropy in -NIVS when applying a momentum flux. And later sections will -demonstrate that this can improve the performance in shear viscosity -simulations. +Compared to the previous NIVS method, the SS-RNEMD approach is +computationally simpler in that only quadratic equations are involved +to determine a set of scaling coefficients, while the NIVS method +required solution of quartic equations. Furthermore, this method +implements {\it isotropic} scaling of velocities in respective slabs, +unlike NIVS, which required extra flexibility in the choice of scaling +coefficients to allow for the energy and momentum constraints. Most +importantly, separating the linear momentum flux from velocity scaling +avoids the underlying cause of the thermal anisotropy in NIVS. In +later sections, we demonstrate that this can improve the calculated +shear viscosities from RNEMD simulations. -This approach is advantageous over the original momentum swapping in -many aspects. In one swapping, the velocity vectors involved are -usually very different (or the generated flux is trivial to obtain -gradients), thus the swapping tends to incur perturbations to the -neighbors of the particles involved. Comparatively, our approach -disperse the flux to every selected particle in a slab so that -perturbations in the flux generating region could be -minimized. Additionally, because the momentum swapping steps tend to -result in a nonthermal distribution, when an imposed flux is -relatively large and diffusions from the neighboring slabs could no -longer remedy this effect, problematic distributions would be -observed. In comparison, the operations of our approach has the nature -of preserving the equilibrium velocity distributions (commonly -Maxwell-Boltzmann), and results in later section will illustrate that -this is helpful to retain thermal distributions in a simulation. +The SS-RNEMD approach has advantages over the original momentum +swapping in many respects. In the swapping method, the velocity +vectors involved are usually very different (or the generated flux +would be quite small), thus the swapping tends to create strong +perturbations in the neighborhood of the particles involved. In +comparison, the SS approach distributes the flux widely to all +particles in a slab so that perturbations in the flux generating +region are minimized. Additionally, momentum swapping steps tend to +result in nonthermal velocity distributions when the imposed flux is +large and diffusion from the neighboring slabs cannot carry momentum +away quickly enough. In comparison, the scaling and shearing moves +made in the SS approach preserve the shapes of the equilibrium +velocity disributions (e.g. Maxwell-Boltzmann). The results presented +in later sections will illustrate that this is quite helpful in +retaining reasonable thermal distributions in a simulation. \section{Computational Details} The algorithm has been implemented in our MD simulation code, -OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with -previous RNEMD methods or equilibrium MD (EMD) methods in homogeneous fluids -(Lennard-Jones and SPC/E water). And taking advantage of the method, -we simulate the interfacial friction of different heterogeneous -interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid -water). +OpenMD.\cite{Meineke:2005gd,openmd} We will first compare the method +with previous RNEMD methods and equilibrium MD in homogeneous +fluids (Lennard-Jones and SPC/E water). We have also used the new +method to simulate the interfacial friction of different heterogeneous +interfaces (Au (111) with organic solvents, Au(111) with SPC/E water, +and the Ice Ih - liquid water interface). \subsection{Simulation Protocols} -The systems to be investigated are set up in orthorhombic simulation +The systems we investigated were set up in orthorhombic simulation cells with periodic boundary conditions in all three dimensions. The -$z$ axis of these cells were longer and set as the temperature and/or -momentum gradient axis. And the cells were evenly divided into $N$ -slabs along this axis, with various $N$ depending on individual -system. The $x$ and $y$ axis were of the same length in homogeneous -systems or had length scale close to each other where heterogeneous -interfaces presents. In all cases, before introducing a nonequilibrium -method to establish steady thermal and/or momentum gradients for -further measurements and calculations, canonical ensemble with a -Nos\'e-Hoover thermostat\cite{hoover85} and microcanonical ensemble -equilibrations were used before data collections. For SPC/E water -simulations, isobaric-isothermal equilibrations\cite{melchionna93} are -performed before the above to reach normal pressure (1 bar); for -interfacial systems, similar equilibrations are used to relax the -surface tensions of the $xy$ plane. +$z$-axes of these cells were typically quite long and served as the +temperature and/or momentum gradient axes. The cells were evenly +divided into $N$ slabs along this axis, with $N$ varying for the +individual system. The $x$ and $y$ axes were of similar lengths in all +simulations. In all cases, before introducing a nonequilibrium method +to establish steady thermal and/or momentum gradients, equilibration +simulations were run under the canonical ensemble with a Nos\'e-Hoover +thermostat\cite{hoover85} followed by further equilibration using +standard constant energy (NVE) conditions. For SPC/E water +simulations, isobaric-isothermal equilibrations\cite{melchionna93} +were performed before equilibration to reach standard densities at +atmospheric pressure (1 bar); for interfacial systems, similar +equilibrations with anisotropic box relaxations are used to relax the +surface tension in the $xy$ plane. -While homogeneous fluid systems can be set up with rather random -configurations, our interfacial systems needs a series of steps to -ensure the interfaces be established properly for computations. The -preparation and equilibration of butanethiol covered gold (111) -surface and further solvation and equilibration process is described -in details as in reference \cite{kuang:AuThl}. +While homogeneous fluid systems can be set up with random +configurations, interfacial systems are typically prepared with a +single crystal face presented perpendicular to the $z$-axis. This +crystal face is aligned in the x-y plane of the periodic cell, and the +solvent occupies the region directly above and below a crystalline +slab. The preparation and equilibration of butanethiol covered +Au(111) surfaces, as well as the solvation and equilibration processes +used for these interfaces are described in detail in reference +\cite{kuang:AuThl}. -As for the ice/liquid water interfaces, the basal surface of ice -lattice was first constructed. Hirsch {\it et - al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice -lattices with all possible proton order configurations. We refer to -their results and choose the configuration of the lowest energy after -geometry optimization as the unit cell for our ice lattices. Although +For the ice / liquid water interfaces, the basal surface of ice +lattice was first constructed. Hirsch and Ojam\"{a}e +\cite{doi:10.1021/jp048434u} explored the energetics of ice lattices +with all possible proton ordered configurations. We utilized Hirsch +and Ojam\"{a}e's structure 6 ($P2_12_12_1$) which is an orthorhombic +cell giving a proton-ordered version of Ice Ih. The basal face of ice +in this unit cell geometry is the $\{0~0~1\}$ face. Although experimental solid/liquid coexistant temperature under normal pressure -should be close to 273K, Bryk and Haymet's simulations of ice/liquid -water interfaces with different models suggest that for SPC/E, the -most stable interface is observed at 225$\pm$5K.\cite{bryk:10258} -Therefore, our ice/liquid water simulations were carried out at -225K. To have extra protection of the ice lattice during initial -equilibration (when the randomly generated liquid phase configuration -could release large amount of energy in relaxation), restraints were -applied to the ice lattice to avoid inadvertent melting by the heat -dissipated from the high enery configurations. -[MAY ADD A SNAPSHOT FOR BASAL PLANE] +are close to 273K, Bryk and Haymet's simulations of ice/liquid water +interfaces with different models suggest that for SPC/E, the most +stable interface is observed at 225$\pm$5K.\cite{bryk:10258} +Therefore, our ice/liquid water simulations were carried out at an +average temperature of 225K. Molecular translation and orientational +restraints were applied in the early stages of equilibration to +prevent melting of the ice slab. These restraints were removed during +NVT equilibration, well before data collection was carried out. \subsection{Force Field Parameters} -For comparison of our new method with previous work, we retain our +For comparing the SS-RNEMD method with previous work, we retained force field parameters consistent with previous simulations. Argon is the Lennard-Jones fluid used here, and its results are reported in -reduced unit for direct comparison purpose. +reduced units for purposes of direct comparison with previous +calculations. -As for our water simulations, SPC/E model is used throughout this work -for consistency. Previous work for transport properties of SPC/E water -model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so -that unnecessary repetition of previous methods can be avoided. +For our water simulations, we utilized the SPC/E model throughout this +work. Previous work for transport properties of SPC/E water model is +available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct +comparison with previous calculation methods is possible. The Au-Au interaction parameters in all simulations are described by the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The @@ -310,36 +348,38 @@ butanethiol and liquid hexane and toluene. The United- For our gold/organic liquid interfaces, the small organic molecules included in our simulations are the Au surface capping agent -butanethiol and liquid hexane and toluene. The United-Atom +butanethiol as well as liquid hexane and liquid toluene. The +United-Atom models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} -for these components were used in this work for better computational +for these components were used in this work for computational efficiency, while maintaining good accuracy. We refer readers to our previous work\cite{kuang:AuThl} for further details of these models, as well as the interactions between Au and the above organic molecule components. \subsection{Thermal conductivities} -When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to -impose kinetic energy transfer, the method can be used for thermal -conductivity computations. Similar to previous RNEMD methods, we -assume linear response of the temperature gradient with respect to the -thermal flux in general case. And the thermal conductivity ($\lambda$) -can be obtained with the imposed kinetic energy flux and the measured +When the linear momentum flux $\vec{j}_z(\vec{p})$ is set to zero and +the target $J_z$ is non-zero, SS-RNEMD imposes kinetic energy transfer +between the slabs, which can be used for computation of thermal +conductivities. Similar to previous RNEMD methods, we assume that we +are in the linear response regime of the temperature gradient with +respect to the thermal flux. The thermal conductivity ($\lambda$) can +be calculated using the imposed kinetic energy flux and the measured thermal gradient: \begin{equation} J_z = -\lambda \frac{\partial T}{\partial z} \end{equation} Like other imposed-flux methods, the energy flux was calculated using the total non-physical energy transferred (${E_{total}}$) from slab -``c'' to slab ``h'', which is recorded throughout a simulation, and -the time for data collection $t$: +``c'' to slab ``h'', which was recorded throughout the simulation, and +the total data collection time $t$: \begin{equation} -J_z = \frac{E_{total}}{2 t L_x L_y} +J_z = \frac{E_{total}}{2 t L_x L_y}. \end{equation} -where $L_x$ and $L_y$ denotes the dimensions of the plane in a +Here, $L_x$ and $L_y$ denote the dimensions of the plane in a simulation cell perpendicular to the thermal gradient, and a factor of -two in the denominator is necessary for the heat transport occurs in -both $+z$ and $-z$ directions. The average temperature gradient +two in the denominator is necessary as the heat transport occurs in +both the $+z$ and $-z$ directions. The average temperature gradient ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear regression of the temperature profile, which is recorded during a simulation for each slab in a cell. For Lennard-Jones simulations, @@ -347,20 +387,18 @@ Alternatively, the method can carry out shear viscosit (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$). \subsection{Shear viscosities} -Alternatively, the method can carry out shear viscosity calculations -by specify a momentum flux. In our algorithm, one can specify the -three components of the flux vector $\vec{j}_z(\vec{p})$ -respectively. For shear viscosity simulations, $j_z(p_z)$ is usually -set to zero. For isotropic systems, the direction of -$\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the -ability of arbitarily specifying the vector direction in our method -could provide convenience in anisotropic simulations. +Alternatively, when the linear momentum flux $\vec{j}_z(\vec{p})$ is +non-zero in either the $x$ or $y$ directions, the method can be used +to compute the shear viscosity. For isotropic systems, the direction +of $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the +ability to arbitarily specify the vector direction in our method could +provide convenience when working with anisotropic interfaces. -Similar to thermal conductivity computations, for a homogeneous -system, linear response of the momentum gradient with respect to the -shear stress is assumed, and the shear viscosity ($\eta$) can be -obtained with the imposed momentum flux (e.g. in $x$ direction) and -the measured gradient: +In a manner similar to the thermal conductivity calculations, a linear +response of the momentum gradient with respect to the shear stress is +assumed, and the shear viscosity ($\eta$) can be obtained with the +imposed momentum flux (e.g. in $x$ direction) and the measured +velocity gradient: \begin{equation} j_z(p_x) = -\eta \frac{\partial v_x}{\partial z} \end{equation} @@ -376,74 +414,78 @@ Although $J_z$ may be switched off for shear viscosity viscosities are also reported in reduced units (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$). -Although $J_z$ may be switched off for shear viscosity simulations at -a certain temperature, our method's ability to impose both a thermal -and a momentum flux in one simulation allows the combination of a -temperature and a velocity gradient. In this case, since viscosity is -generally a function of temperature, the local viscosity also depends -on the local temperature. Therefore, in one such simulation, viscosity -at $z$ (corresponding to a certain $T$) can be computed with the -applied shear flux and the local velocity gradient (which can be -obtained by finite difference approximation). As a whole, the -viscosity can be mapped out as the function of temperature in one -single trajectory of simulation. Results for shear viscosity -computations of SPC/E water will demonstrate its effectiveness in -detail. +Although $J_z$ may be switched off for shear viscosity simulations, +the SS-RNEMD method allows the user the ability to simultaneously +impose both a thermal and a momentum flux during a single +simulation. This can create system with coincident temperature and a +velocity gradients. Since the viscosity is generally a function of +temperature, the local viscosity depends on the local temperature in +the fluid. Therefore, in a single simulation, viscosity at $z$ +(corresponding to a certain $T$) can be computed with the applied +shear flux and the local velocity gradient (which can be obtained by +finite difference approximation). This means that the temperature +dependence of the viscosity can be mapped out in only one +trajectory. Results for shear viscosity computations of SPC/E water +will demonstrate SS-RNEMD's efficiency in this respect. -\subsection{Interfacial friction and Slip length} -While the shear stress results in a velocity gradient within bulk -fluid phase, its effect at a solid-liquid interface could vary due to -the interaction strength between the two phases. The interfacial -friction coefficient $\kappa$ is defined to relate the shear stress -(e.g. along $x$-axis) with the relative fluid velocity tangent to the -interface: +\subsection{Interfacial friction and slip length} +Shear stress creates a velocity gradient within bulk fluid phases, but +at solid-liquid interfaces, the effects of a shear stress depend on +the molecular details of the interface. The interfacial friction +coefficient, $\kappa$, relates the shear stress (e.g. along the +$x$-axis) with the relative fluid velocity tangent to the interface: \begin{equation} -j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface} +j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right] \end{equation} -Under ``stick'' boundary condition, $\Delta v_x|_{interface} -\rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for -``slip'' boundary conditions at the solid-liquid interfaces, $\kappa$ -becomes finite. To characterize the interfacial boundary conditions, -slip length ($\delta$) is defined using $\kappa$ and the shear -viscocity of liquid phase ($\eta$): +where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured +directly adjacent to the interface. Under ``stick'' boundary +condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads +to $\kappa\rightarrow\infty$. However, for ``slip'' boundary +conditions at a solid-liquid interface, $\kappa$ becomes finite. To +characterize the interfacial boundary conditions, the slip length, +$\delta$, is defined by the ratio of the fluid-phase viscosity to the +friction coefficient of the interface: \begin{equation} \delta = \frac{\eta}{\kappa} \end{equation} -so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition, -and depicts how ``slippery'' an interface is. Figure \ref{slipLength} -illustrates how this quantity is defined and computed for a -solid-liquid interface. [MAY INCLUDE SNAPSHOT IN FIGURE] +In ``no-slip'' or ``stick'' boundary conditions, $\delta\rightarrow +0$, and $\delta$ is a measure of how ``slippery'' an interface is. +Figure \ref{slipLength} illustrates how this quantity is defined and +computed for a solid-liquid interface. \begin{figure} \includegraphics[width=\linewidth]{defDelta} \caption{The slip length $\delta$ can be obtained from a velocity profile of a solid-liquid interface simulation, when a momentum flux - is applied. An example of Au/hexane interfaces is shown, and the - calculation for the left side is illustrated. The calculation for - the right side is similar to the left.} + is applied. The data shown is for a simulated Au/hexane interface. + The Au crystalline region is moving as a block (lower dots), while + the measured velocity gradient in the hexane phase is discontinuous + a the interface.} \label{slipLength} \end{figure} -In our method, a shear stress can be applied similar to shear -viscosity computations by applying an unphysical momentum flux -(e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as -shown in Figure \ref{slipLength}, in which the velocity gradients -within liquid phase and velocity difference at the liquid-solid -interface can be measured respectively. Further calculations and -characterizations of the interface can be carried out using these -data. +Since the method can be applied for interfaces as well as for bulk +materials, the shear stress is applied in an identical manner to the +shear viscosity computations, e.g. by applying an unphysical momentum +flux, $j_z(\vec{p})$. With the correct choice of $\vec{p}$ in the +$x-y$ plane, one can compute friction coefficients and slip lengths +for a number of different dragging vectors on a given slab. The +corresponding velocity profiles can be obtained as shown in Figure +\ref{slipLength}, in which the velocity gradients within the liquid +phase and the velocity difference at the liquid-solid interface can be +easily measured from saved simulation data. \section{Results and Discussions} \subsection{Lennard-Jones fluid} -Our orthorhombic simulation cell of Lennard-Jones fluid has identical -parameters to our previous work\cite{kuang:164101} to facilitate -comparison. Thermal conductivitis and shear viscosities were computed -with the algorithm applied to the simulations. The results of thermal -conductivity are compared with our previous NIVS algorithm. However, -since the NIVS algorithm could produce temperature anisotropy for -shear viscocity computations, these results are instead compared to -the momentum swapping approaches. Table \ref{LJ} lists these -calculations with various fluxes in reduced units. +Our orthorhombic simulation cell for the Lennard-Jones fluid has +identical parameters to our previous work\cite{kuang:164101} to +facilitate comparison. Thermal conductivities and shear viscosities +were computed with the new algorithm applied to the simulations. The +results of thermal conductivity are compared with our previous NIVS +algorithm. However, since the NIVS algorithm produced temperature +anisotropy for shear viscocity computations, these results are instead +compared to the previous momentum swapping approaches. Table \ref{LJ} +lists these values with various fluxes in reduced units. \begin{table*} \begin{minipage}{\linewidth} @@ -462,7 +504,7 @@ calculations with various fluxes in reduced units. \multicolumn{2}{c}{$\lambda^*$} & \multicolumn{2}{c}{$\eta^*$} \\ \hline - Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ & + Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ & NIVS\cite{kuang:164101} & This work & Swapping & This work \\ \hline 0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\ @@ -498,7 +540,7 @@ applying a thermal flux). In this sense, this method a or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha P^\alpha$) would not achieve this effect unless thermal flux vanishes (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which do not contribute to -applying a thermal flux). In this sense, this method aids to achieve +applying a thermal flux). In this sense, this method aids in achieving minimal perturbation to a simulation while imposing a thermal flux. \subsubsection{Shear viscosity}