ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
Revision: 3652
Committed: Mon Oct 18 18:27:24 2010 UTC (13 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 20608 byte(s)
Log Message:
editing the methodology & intro sections

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{setspace}
5 \usepackage{endfloat}
6 \usepackage{caption}
7 \usepackage{graphicx}
8 \usepackage{multirow}
9 \usepackage[square, comma, sort&compress]{natbib}
10 \usepackage{url}
11 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
12 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
13 9.0in \textwidth 6.5in \brokenpenalty=10000
14
15 % double space list of tables and figures
16 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
17 \setlength{\abovecaptionskip}{20 pt}
18 \setlength{\belowcaptionskip}{30 pt}
19
20 \bibpunct{[}{]}{,}{s}{}{;}
21 \bibliographystyle{aip}
22
23 \begin{document}
24
25 \title{The Langevin Hull: Constant pressure and temperature dynamics for non-periodic systems}
26
27 \author{Charles F. Vardeman II, Kelsey M. Stocker, and J. Daniel
28 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
29 Department of Chemistry and Biochemistry,\\
30 University of Notre Dame\\
31 Notre Dame, Indiana 46556}
32
33 \date{\today}
34
35 \maketitle
36
37 \begin{doublespace}
38
39 \begin{abstract}
40 We have developed a new isobaric-isothermal (NPT) algorithm which
41 applies an external pressure to the facets comprising the convex
42 hull surrounding the objects in the system. Additionally, a Langevin
43 thermostat is applied to facets of the hull to mimic contact with an
44 external heat bath. This new method, the ``Langevin Hull'', performs
45 better than traditional affine transform methods for systems
46 containing heterogeneous mixtures of materials with different
47 compressibilities. It does not suffer from the edge effects of
48 boundary potential methods, and allows realistic treatment of both
49 external pressure and thermal conductivity to an implicit solvent.
50 We apply this method to several different systems including bare
51 nanoparticles, nanoparticles in an explicit solvent, as well as
52 clusters of liquid water and ice. The predicted mechanical and
53 thermal properties of these systems are in good agreement with
54 experimental data.
55 \end{abstract}
56
57 \newpage
58
59 %\narrowtext
60
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 % BODY OF TEXT
63 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64
65
66 \section{Introduction}
67
68 The most common molecular dynamics methods for sampling configurations
69 of an isobaric-isothermal (NPT) ensemble attempt to maintain a target
70 pressure in a simulation by coupling the volume of the system to an
71 extra degree of freedom, the {\it barostat}. These methods require
72 periodic boundary conditions, because when the instantaneous pressure
73 in the system differs from the target pressure, the volume is
74 typically reduced or expanded using {\it affine transforms} of the
75 system geometry. An affine transform scales both the box lengths as
76 well as the scaled particle positions (but not the sizes of the
77 particles). The most common constant pressure methods, including the
78 Melchionna modification\cite{Melchionna1993} to the
79 Nos\'e-Hoover-Andersen equations of
80 motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx} the Berendsen
81 pressure bath,\cite{ISI:A1984TQ73500045} and the Langevin
82 Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize coordinate
83 transformation to adjust the box volume.
84
85 As long as the material in the simulation box is essentially a bulk
86 liquid which has a relatively uniform compressibility, the standard
87 approach provides an excellent way of adjusting the volume of the
88 system and applying pressure directly via the interactions between
89 atomic sites.
90
91 The problem with these approaches becomes apparent when the material
92 being simulated is an inhomogeneous mixture in which portions of the
93 simulation box are incompressible relative to other portions.
94 Examples include simulations of metallic nanoparticles in liquid
95 environments, proteins at interfaces, as well as other multi-phase or
96 interfacial environments. In these cases, the affine transform of
97 atomic coordinates will either cause numerical instability when the
98 sites in the incompressible medium collide with each other, or lead to
99 inefficient sampling of system volumes if the barostat is set slow
100 enough to avoid collisions in the incompressible region.
101
102 \begin{figure}
103 \includegraphics[width=\linewidth]{AffineScale2}
104 \caption{Affine Scaling constant pressure methods use box-length
105 scaling to adjust the volume to adjust to under- or over-pressure
106 conditions. In a system with a uniform compressibility (e.g. bulk
107 fluids) these methods can work well. In systems containing
108 heterogeneous mixtures, the affine scaling moves required to adjust
109 the pressure in the high-compressibility regions can cause molecules
110 in low compressibility regions to collide.}
111 \label{affineScale}
112 \end{figure}
113
114 Additionally, one may often wish to simulate explicitly non-periodic
115 systems, and the constraint that a periodic box must be used to
116
117 Explicitly non-periodic systems
118
119 Elastic Bag
120
121 Spherical Boundary approaches
122
123 \section{Methodology}
124
125 We have developed a new method which uses a constant pressure and
126 temperature bath. This bath interacts only with the objects that are
127 currently at the edge of the system. Since the edge is determined
128 dynamically as the simulation progresses, no {\it a priori} geometry
129 is defined. The pressure and temperature bath interacts {\it
130 directly} with the atoms on the edge and not with atoms interior to
131 the simulation. This means that there are no affine transforms
132 required. There are also no fictitious particles or bounding
133 potentials used in this approach.
134
135 The basics of the method are as follows. The simulation starts as a
136 collection of atomic locations in three dimensions (a point cloud).
137 Delaunay triangulation is used to find all facets between coplanar
138 neighbors. In highly symmetric point clouds, facets can contain many
139 atoms, but in all but the most symmetric of cases one might experience
140 in a molecular dynamics simulation, the facets are simple triangles in
141 3-space that contain exactly three atoms.
142
143 The convex hull is the set of facets that have {\it no concave
144 corners} at an atomic site. This eliminates all facets on the
145 interior of the point cloud, leaving only those exposed to the
146 bath. Sites on the convex hull are dynamic. As molecules re-enter the
147 cluster, all interactions between atoms on that molecule and the
148 external bath are removed.
149
150 For atomic sites in the interior of the point cloud, the equations of
151 motion are simple Newtonian dynamics,
152 \begin{equation}
153 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
154 \label{eq:Newton}
155 \end{equation}
156 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
157 instantaneous velocity of site $i$ at time $t$, and $U$ is the total
158 potential energy. For atoms on the exterior of the cluster
159 (i.e. those that occupy one of the vertices of the convex hull), the
160 equation of motion is modified with an external force, ${\mathbf
161 F}_i^{\mathrm ext}$,
162 \begin{equation}
163 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
164 \end{equation}
165
166 The external bath interacts directly with the facets of the convex
167 hull. Since each vertex (or atom) provides one corner of a triangular
168 facet, the force on the facets are divided equally to each vertex.
169 However, each vertex can participate in multiple facets, so the resultant
170 force is a sum over all facets $f$ containing vertex $i$:
171 \begin{equation}
172 {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
173 } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
174 F}_f^{\mathrm ext}
175 \end{equation}
176
177 The external pressure bath applies a force to the facets of the convex
178 hull in direct proportion to the area of the facet, while the thermal
179 coupling depends on the solvent temperature, friction and the size and
180 shape of each facet. The thermal interactions are expressed as a
181 typical Langevin description of the forces,
182 \begin{equation}
183 \begin{array}{rclclcl}
184 {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
185 & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
186 \end{array}
187 \end{equation}
188 Here, $P$ is the external pressure, $A_f$ and $\hat{n}_f$ are the area
189 and normal vectors for facet $f$, respectively. ${\mathbf v}_f(t)$ is
190 the velocity of the facet,
191 \begin{equation}
192 {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
193 \end{equation}
194 and $\Xi_f(t)$ is a ($3 \times 3$) hydrodynamic tensor that depends on
195 the geometry and surface area of facet $f$ and the viscosity of the
196 fluid (See Appendix A). The hydrodynamic tensor is related to the
197 fluctuations of the random force, $\mathbf{R}(t)$, by the
198 fluctuation-dissipation theorem,
199 \begin{eqnarray}
200 \left< {\mathbf R}_f(t) \right> & = & 0 \\
201 \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
202 \Xi_f(t)\delta(t-t^\prime).
203 \label{eq:randomForce}
204 \end{eqnarray}
205
206 Once the hydrodynamic tensor is known for a given facet (see Appendix
207 A) obtaining a stochastic vector that has the properties in
208 Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
209 one-time Cholesky decomposition to obtain the square root matrix of
210 the resistance tensor,
211 \begin{equation}
212 \Xi_f = {\bf S} {\bf S}^{T},
213 \label{eq:Cholesky}
214 \end{equation}
215 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
216 vector with the statistics required for the random force can then be
217 obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
218 has elements chosen from a Gaussian distribution, such that:
219 \begin{equation}
220 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
221 {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
222 \end{equation}
223 where $\delta t$ is the timestep in use during the simulation. The
224 random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
225 have the correct properties required by Eq. (\ref{eq:randomForce}).
226
227 We have implemented this method by extending the Langevin dynamics
228 integrator in our group code, OpenMD.\cite{Meineke2005,openmd}
229
230 \section{Tests \& Applications}
231
232 \subsection{Bulk modulus of gold nanoparticles}
233
234 \begin{figure}
235 \includegraphics[width=\linewidth]{pressure_tb}
236 \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
237 pressure = 4 GPa}
238 \label{pressureResponse}
239 \end{figure}
240
241 \begin{figure}
242 \includegraphics[width=\linewidth]{temperature_tb}
243 \caption{Temperature equilibration depends on surface area and bath
244 viscosity. Target Temperature = 300K}
245 \label{temperatureResponse}
246 \end{figure}
247
248 \begin{equation}
249 \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
250 P}\right)
251 \end{equation}
252
253 \begin{figure}
254 \includegraphics[width=\linewidth]{compress_tb}
255 \caption{Isothermal Compressibility (18 \AA gold nanoparticle)}
256 \label{temperatureResponse}
257 \end{figure}
258
259 \subsection{Compressibility of SPC/E water clusters}
260
261 Both NVT \cite{Glattli2002} and NPT \cite{Motakabbir1990, Pi2009} molecular dynamics simulations of SPC/E water have yielded values for the isothermal compressibility of water that agree well with experiment \cite{Fine1973}. The results of three different methods for computing the isothermal compressibility from Langevin Hull simulations for pressures between 1 and 6500 atm are shown in Fig. 5 along with compressibility values obtained from both other SPC/E simulations and experiment. Compressibility values from all references are for applied pressures within the range 1 - 1000 atm.
262
263 \begin{figure}
264 \includegraphics[width=\linewidth]{new_isothermal}
265 \caption{Compressibility of SPC/E water}
266 \label{compWater}
267 \end{figure}
268
269 We initially used the classic compressibility formula
270
271 \begin{equation}
272 \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right )_{T}
273 \end{equation}
274
275 to calculate the the isothermal compressibility at each target pressure. These calculations yielded compressibility values that were dramatically higher than both previous simulations and experiment. The particular compressibility expression used requires the calculation of both a volume and pressure differential, thereby stipulating that the data from at least two simulations at different pressures must be used to calculate the isothermal compressibility at one pressure.
276
277 Per the fluctuation dissipation theorem \cite{Debenedetti1986}, the hull volume fluctuation in any given simulation can be used to calculated the isothermal compressibility at that particular pressure
278
279 \begin{equation}
280 \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle V \right \rangle ^{2}}{V \, k_{B} \, T}
281 \end{equation}
282
283 Thus, the compressibility of each simulation run can be calculated entirely independently from all other trajectories. However, the resulting compressibilities were still as much as an order of magnitude larger than the reference values. The effect was particularly pronounced at the low end of the pressure range. At ambient temperature and low pressures, there exists an equilibrium between vapor and liquid phases. Vapor molecules are naturally more diffuse around the exterior of the cluster, causing artificially large cluster volumes. Any compressibility calculation that relies on the hull volume will suffer these effects.
284
285 In order to calculate the isothermal compressibility without being hindered by hull volume issues, we adapted the classic compressibility formula so that the compressibility could be calculated using information about the local density instead of the volume of the convex hull. We calculated the $g_{OO}(r)$ for a 1 nanosecond simulation of a cluster of 1372 SPC/E water molecules and spherically integrated the function over the bounds 0 to $r'$. In all cases, the value of $r'$ was 17.26216 $\AA$. The value of the total integral between these bounds is essentially the number (N) of molecules within volume $\frac{4}{3}\pi r'^{3}$ at a given pressure. To yield an actual molecule count, N must be scaled by an ideal density. However, even in the absence of an ideal density, we can use the relationship $\rho = \frac{N}{V}$ to rewrite the isothermal compressibility formula as
286
287 \begin{equation}
288 \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right )_{T}
289 \end{equation}
290
291 Isothermal compressibility values calculated using this modified expression are in good agreement with the reference values throughout the 1 - 1000 atm pressure regime. Regardless of the difficulty in obtaining accurate hull volumes at low temperature and pressures, the Langevin Hull NPT method provides reasonable isothermal compressibility values for water through a large range of pressures.
292
293 \subsection{Molecular orientation distribution at cluster boundary}
294
295 In order for non-periodic boundary conditions to be widely applicable, they must be constructed in such a way that they allow a finite, usually small, simulated system to replicate the properties of an infinite bulk system. Naturally, this requirement has spawned many methods for inserting boundaries into simulated systems [REF... ?]. Of particular interest to our characterization of the Langevin Hull is the orientation of water molecules included in the geometric hull. Ideally, all molecules in the cluster will have the same orientational distribution as bulk water.
296
297 The orientation of molecules at the edges of a simulated cluster has long been a concern when performing simulations of explicitly non-periodic systems. Early work led to the surface constrained soft sphere dipole model (SCSSD) \cite{Warshel1978} in which the surface molecules are fixed in a random orientation representative of the bulk solvent structural properties. Belch, et al \cite{Belch1985} simulated clusters of TIPS2 water surrounded by a hydrophobic bounding potential. The spherical hydrophobic boundary induced dangling hydrogen bonds at the surface that propagated deep into the cluster, affecting 70\% of the 100 molecules in the simulation. This result echoes an earlier study which showed that an extended planar hydrophobic surface caused orientational preference at the surface which extended 7 \r{A} into the liquid simulation cell \cite{Lee1984}. The surface constrained all-atom solvent (SCAAS) model \cite{King1989} improved upon its SCSSD predecessor. The SCAAS model utilizes a polarization constraint which is applied to the surface molecules to maintain bulk-like structure at the cluster surface. A radial constraint is used to maintain the desired bulk density of the liquid. Both constraint forces are applied only to a pre-determined number of the outermost molecules.
298
299 In contrast, the Langevin Hull does not require that the orientation of molecules be fixed, nor does it utilize an explicitly hydrophobic boundary, orientational constraint or radial constraint. The number and identity of the molecules included on the convex hull are dynamic properties, thus avoiding the formation of an artificial solvent boundary layer. The hope is that the water molecules on the surface of the cluster, if left to their own devices in the absence of orientational and radial constraints, will maintain a bulk-like orientational distribution.
300
301 To determine the extent of these effects demonstrated by the Langevin Hull, we examined the orientations exhibited by SPC/E water in a cluster of 1372 molecules at 300 K and at pressures ranging from 1 - 1000 atm.
302
303 The orientation of a water molecule is described by
304
305 \begin{equation}
306 \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
307 \end{equation}
308
309 where $\vec{r}_{i}$ is the vector between molecule {\it i}'s center of mass and the cluster center of mass and $\vec{\mu}_{i}$ is the vector bisecting the H-O-H angle of molecule {\it i}.
310
311 \begin{figure}
312 \includegraphics[width=\linewidth]{g_r_theta}
313 \caption{Definition of coordinates}
314 \label{coords}
315 \end{figure}
316
317 Fig. 7 shows the probability of each value of $\cos{\theta}$ for molecules in the interior of the cluster (squares) and for molecules included in the convex hull (circles).
318
319 \begin{figure}
320 \includegraphics[width=\linewidth]{pAngle}
321 \caption{SPC/E water clusters: only minor dewetting at the boundary}
322 \label{pAngle}
323 \end{figure}
324
325 As expected, interior molecules (those not included in the convex hull) maintain a bulk-like structure with a uniform distribution of orientations. Molecules included in the convex hull show a slight preference for values of $\cos{\theta} < 0.$ These values correspond to molecules with a hydrogen directed toward the exterior of the cluster, forming a dangling hydrogen bond.
326
327 In the absence of an electrostatic contribution from the exterior bath, the orientational distribution of water molecules included in the Langevin Hull will slightly resemble the distribution at a neat water liquid/vapor interface. Previous molecular dynamics simulations of SPC/E water \cite{Taylor1996} have shown that molecules at the liquid/vapor interface favor an orientation where one hydrogen protrudes from the liquid phase. This behavior is demonstrated by experiments \cite{Du1994} \cite{Scatena2001} showing that approximately one-quarter of water molecules at the liquid/vapor interface form dangling hydrogen bonds. The negligible preference shown in these cluster simulations could be removed through the introduction of an implicit solvent model, which would provide the missing electrostatic interactions between the cluster molecules and the surrounding temperature/pressure bath.
328
329 The orientational preference exhibited by hull molecules is significantly weaker than the preference caused by an explicit hydrophobic bounding potential. Additionally, the Langevin Hull does not require that the orientation of any molecules be fixed in order to maintain bulk-like structure, even at the cluster surface.
330
331
332 \subsection{Heterogeneous nanoparticle / water mixtures}
333
334
335 \section{Appendix A: Hydrodynamic tensor for triangular facets}
336
337 \begin{figure}
338 \includegraphics[width=\linewidth]{hydro}
339 \caption{Hydro}
340 \label{hydro}
341 \end{figure}
342
343 \begin{equation}
344 \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}
345 \end{equation}
346
347 \begin{equation}
348 T_{if}=\frac{A_i}{8\pi\eta R_{if}}\left(I +
349 \frac{\mathbf{R}_{if}\mathbf{R}_{if}^T}{R_{if}^2}\right)
350 \end{equation}
351
352 \section{Appendix B: Computing Convex Hulls on Parallel Computers}
353
354 \section{Acknowledgments}
355 Support for this project was provided by the
356 National Science Foundation under grant CHE-0848243. Computational
357 time was provided by the Center for Research Computing (CRC) at the
358 University of Notre Dame.
359
360 \newpage
361
362 \bibliography{langevinHull}
363
364 \end{doublespace}
365 \end{document}