ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
Revision: 3723
Committed: Fri Feb 18 14:08:15 2011 UTC (13 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 43474 byte(s)
Log Message:
final version

File Contents

# User Rev Content
1 gezelter 3640 \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 kstocke1 3716 \bibpunct{}{}{,}{s}{}{;}
21 gezelter 3667 \bibliographystyle{achemso}
22 gezelter 3640
23     \begin{document}
24    
25     \title{The Langevin Hull: Constant pressure and temperature dynamics for non-periodic systems}
26    
27 kstocke1 3644 \author{Charles F. Vardeman II, Kelsey M. Stocker, and J. Daniel
28 gezelter 3640 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 gezelter 3665 hull surrounding the system. A Langevin thermostat is also applied
43 gezelter 3684 to the facets to mimic contact with an external heat bath. This new
44     method, the ``Langevin Hull'', can handle heterogeneous mixtures of
45 gezelter 3722 materials with different compressibilities. These systems are
46     problematic for traditional affine transform methods. The Langevin
47     Hull does not suffer from the edge effects of boundary potential
48     methods, and allows realistic treatment of both external pressure
49     and thermal conductivity due to the presence of an implicit solvent.
50     We apply this method to several different systems including bare
51     metal nanoparticles, nanoparticles in an explicit solvent, as well
52     as clusters of liquid water. The predicted mechanical properties of
53     these systems are in good agreement with experimental data and
54     previous simulation work.
55 gezelter 3640 \end{abstract}
56    
57     \newpage
58    
59     %\narrowtext
60    
61     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62     % BODY OF TEXT
63     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64    
65    
66     \section{Introduction}
67    
68 gezelter 3641 The most common molecular dynamics methods for sampling configurations
69 gezelter 3667 from an isobaric-isothermal (NPT) ensemble maintain a target pressure
70     in a simulation by coupling the volume of the system to a {\it
71     barostat}, which is an extra degree of freedom propagated along with
72     the particle coordinates. These methods require periodic boundary
73     conditions, because when the instantaneous pressure in the system
74     differs from the target pressure, the volume is reduced or expanded
75     using {\it affine transforms} of the system geometry. An affine
76     transform scales the size and shape of the periodic box as well as the
77     particle positions within the box (but not the sizes of the
78     particles). The most common constant pressure methods, including the
79     Melchionna modification\cite{Melchionna1993} to the
80     Nos\'e-Hoover-Andersen equations of
81     motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx} the Berendsen
82     pressure bath,\cite{ISI:A1984TQ73500045} and the Langevin
83     Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize scaled
84 gezelter 3665 coordinate transformation to adjust the box volume. As long as the
85 gezelter 3667 material in the simulation box has a relatively uniform
86     compressibility, the standard affine transform approach provides an
87     excellent way of adjusting the volume of the system and applying
88     pressure directly via the interactions between atomic sites.
89 gezelter 3652
90 gezelter 3665 One problem with this approach appears when the system being simulated
91     is an inhomogeneous mixture in which portions of the simulation box
92     are incompressible relative to other portions. Examples include
93     simulations of metallic nanoparticles in liquid environments, proteins
94     at ice / water interfaces, as well as other heterogeneous or
95 gezelter 3652 interfacial environments. In these cases, the affine transform of
96     atomic coordinates will either cause numerical instability when the
97 gezelter 3665 sites in the incompressible medium collide with each other, or will
98     lead to inefficient sampling of system volumes if the barostat is set
99     slow enough to avoid the instabilities in the incompressible region.
100 gezelter 3652
101 gezelter 3640 \begin{figure}
102 gezelter 3641 \includegraphics[width=\linewidth]{AffineScale2}
103 gezelter 3667 \caption{Affine scaling methods use box-length scaling to adjust the
104     volume to adjust to under- or over-pressure conditions. In a system
105     with a uniform compressibility (e.g. bulk fluids) these methods can
106     work well. In systems containing heterogeneous mixtures, the affine
107     scaling moves required to adjust the pressure in the
108     high-compressibility regions can cause molecules in low
109     compressibility regions to collide.}
110 gezelter 3640 \label{affineScale}
111     \end{figure}
112    
113 gezelter 3653 One may also wish to avoid affine transform periodic boundary methods
114     to simulate {\it explicitly non-periodic systems} under constant
115     pressure conditions. The use of periodic boxes to enforce a system
116 gezelter 3665 volume requires either effective solute concentrations that are much
117 gezelter 3653 higher than desirable, or unreasonable system sizes to avoid this
118 gezelter 3684 effect. For example, calculations using typical hydration boxes
119 gezelter 3653 solvating a protein under periodic boundary conditions are quite
120 gezelter 3689 expensive. A 62 \AA$^3$ box of water solvating a moderately small
121 gezelter 3684 protein like hen egg white lysozyme (PDB code: 1LYZ) yields an
122     effective protein concentration of 100 mg/mL.\cite{Asthagiri20053300}
123 gezelter 3640
124 kstocke1 3715 {\it Total} protein concentrations in the cell are typically on the
125 gezelter 3689 order of 160-310 mg/ml,\cite{Brown1991195} and individual proteins
126     have concentrations orders of magnitude lower than this in the
127     cellular environment. The effective concentrations of single proteins
128     in simulations may have significant effects on the structure and
129 gezelter 3722 dynamics of simulated systems.
130 gezelter 3684
131 gezelter 3665 \subsection*{Boundary Methods}
132 gezelter 3667 There have been a number of approaches to handle simulations of
133     explicitly non-periodic systems that focus on constant or
134     nearly-constant {\it volume} conditions while maintaining bulk-like
135     behavior. Berkowitz and McCammon introduced a stochastic (Langevin)
136     boundary layer inside a region of fixed molecules which effectively
137     enforces constant temperature and volume (NVT)
138     conditions.\cite{Berkowitz1982} In this approach, the stochastic and
139     fixed regions were defined relative to a central atom. Brooks and
140     Karplus extended this method to include deformable stochastic
141     boundaries.\cite{iii:6312} The stochastic boundary approach has been
142 gezelter 3689 used widely for protein simulations.
143 gezelter 3640
144 gezelter 3653 The electrostatic and dispersive behavior near the boundary has long
145 gezelter 3665 been a cause for concern when performing simulations of explicitly
146     non-periodic systems. Early work led to the surface constrained soft
147     sphere dipole model (SCSSD)\cite{Warshel1978} in which the surface
148     molecules are fixed in a random orientation representative of the bulk
149     solvent structural properties. Belch {\it et al.}\cite{Belch1985}
150     simulated clusters of TIPS2 water surrounded by a hydrophobic bounding
151     potential. The spherical hydrophobic boundary induced dangling
152     hydrogen bonds at the surface that propagated deep into the cluster,
153 gezelter 3689 affecting most of the molecules in the simulation. This result echoes
154     an earlier study which showed that an extended planar hydrophobic
155     surface caused orientational preferences at the surface which extended
156     relatively deep (7 \AA) into the liquid simulation cell.\cite{Lee1984}
157     The surface constrained all-atom solvent (SCAAS) model \cite{King1989}
158     improved upon its SCSSD predecessor. The SCAAS model utilizes a
159     polarization constraint which is applied to the surface molecules to
160     maintain bulk-like structure at the cluster surface. A radial
161     constraint is used to maintain the desired bulk density of the
162     liquid. Both constraint forces are applied only to a pre-determined
163     number of the outermost molecules.
164 gezelter 3640
165 gezelter 3665 Beglov and Roux have developed a boundary model in which the hard
166     sphere boundary has a radius that varies with the instantaneous
167     configuration of the solute (and solvent) molecules.\cite{beglov:9050}
168     This model contains a clear pressure and surface tension contribution
169 gezelter 3689 to the free energy.
170 gezelter 3640
171 gezelter 3665 \subsection*{Restraining Potentials}
172 gezelter 3653 Restraining {\it potentials} introduce repulsive potentials at the
173     surface of a sphere or other geometry. The solute and any explicit
174 gezelter 3665 solvent are therefore restrained inside the range defined by the
175     external potential. Often the potentials include a weak short-range
176     attraction to maintain the correct density at the boundary. Beglov
177     and Roux have also introduced a restraining boundary potential which
178     relaxes dynamically depending on the solute geometry and the force the
179     explicit system exerts on the shell.\cite{Beglov:1995fk}
180 gezelter 3653
181 gezelter 3665 Recently, Krilov {\it et al.} introduced a {\it flexible} boundary
182     model that uses a Lennard-Jones potential between the solvent
183     molecules and a boundary which is determined dynamically from the
184 gezelter 3689 position of the nearest solute atom.\cite{LiY._jp046852t,Zhu:2008fk} This
185 gezelter 3665 approach allows the confining potential to prevent solvent molecules
186     from migrating too far from the solute surface, while providing a weak
187     attractive force pulling the solvent molecules towards a fictitious
188     bulk solvent. Although this approach is appealing and has physical
189     motivation, nanoparticles do not deform far from their original
190     geometries even at temperatures which vaporize the nearby solvent. For
191     the systems like this, the flexible boundary model will be nearly
192 gezelter 3653 identical to a fixed-volume restraining potential.
193    
194 gezelter 3665 \subsection*{Hull methods}
195 gezelter 3653 The approach of Kohanoff, Caro, and Finnis is the most promising of
196     the methods for introducing both constant pressure and temperature
197     into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
198     This method is based on standard Langevin dynamics, but the Brownian
199     or random forces are allowed to act only on peripheral atoms and exert
200 gezelter 3689 forces in a direction that is inward-facing relative to the facets of
201     a closed bounding surface. The statistical distribution of the random
202 gezelter 3653 forces are uniquely tied to the pressure in the external reservoir, so
203     the method can be shown to sample the isobaric-isothermal ensemble.
204     Kohanoff {\it et al.} used a Delaunay tessellation to generate a
205     bounding surface surrounding the outermost atoms in the simulated
206     system. This is not the only possible triangulated outer surface, but
207     guarantees that all of the random forces point inward towards the
208     cluster.
209    
210     In the following sections, we extend and generalize the approach of
211     Kohanoff, Caro, and Finnis. The new method, which we are calling the
212     ``Langevin Hull'' applies the external pressure, Langevin drag, and
213 gezelter 3667 random forces on the {\it facets of the hull} instead of the atomic
214     sites comprising the vertices of the hull. This allows us to decouple
215     the external pressure contribution from the drag and random force.
216     The methodology is introduced in section \ref{sec:meth}, tests on
217     crystalline nanoparticles, liquid clusters, and heterogeneous mixtures
218     are detailed in section \ref{sec:tests}. Section \ref{sec:discussion}
219     summarizes our findings.
220 gezelter 3653
221 gezelter 3640 \section{Methodology}
222 gezelter 3653 \label{sec:meth}
223 gezelter 3640
224 gezelter 3665 The Langevin Hull uses an external bath at a fixed constant pressure
225 gezelter 3689 ($P$) and temperature ($T$) with an effective solvent viscosity
226     ($\eta$). This bath interacts only with the objects on the exterior
227     hull of the system. Defining the hull of the atoms in a simulation is
228     done in a manner similar to the approach of Kohanoff, Caro and
229     Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous configuration
230     of the atoms in the system is considered as a point cloud in three
231     dimensional space. Delaunay triangulation is used to find all facets
232     between coplanar
233     neighbors.\cite{delaunay,springerlink:10.1007/BF00977785} In highly
234 gezelter 3665 symmetric point clouds, facets can contain many atoms, but in all but
235 gezelter 3689 the most symmetric of cases, the facets are simple triangles in
236     3-space which contain exactly three atoms.
237 gezelter 3640
238 gezelter 3652 The convex hull is the set of facets that have {\it no concave
239 gezelter 3665 corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
240     eliminates all facets on the interior of the point cloud, leaving only
241     those exposed to the bath. Sites on the convex hull are dynamic; as
242     molecules re-enter the cluster, all interactions between atoms on that
243     molecule and the external bath are removed. Since the edge is
244     determined dynamically as the simulation progresses, no {\it a priori}
245     geometry is defined. The pressure and temperature bath interacts only
246 gezelter 3660 with the atoms on the edge and not with atoms interior to the
247     simulation.
248 gezelter 3640
249 gezelter 3662 \begin{figure}
250 gezelter 3688 \includegraphics[width=\linewidth]{solvatedNano}
251 gezelter 3662 \caption{The external temperature and pressure bath interacts only
252     with those atoms on the convex hull (grey surface). The hull is
253 gezelter 3667 computed dynamically at each time step, and molecules can move
254 kstocke1 3694 between the interior (Newtonian) region and the Langevin Hull.}
255 gezelter 3662 \label{fig:hullSample}
256     \end{figure}
257    
258 gezelter 3665 Atomic sites in the interior of the simulation move under standard
259 gezelter 3660 Newtonian dynamics,
260 gezelter 3640 \begin{equation}
261 gezelter 3652 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
262     \label{eq:Newton}
263 gezelter 3640 \end{equation}
264 gezelter 3652 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
265     instantaneous velocity of site $i$ at time $t$, and $U$ is the total
266     potential energy. For atoms on the exterior of the cluster
267     (i.e. those that occupy one of the vertices of the convex hull), the
268     equation of motion is modified with an external force, ${\mathbf
269 kstocke1 3695 F}_i^{\mathrm ext}$:
270 gezelter 3640 \begin{equation}
271 gezelter 3652 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
272 gezelter 3640 \end{equation}
273    
274 gezelter 3665 The external bath interacts indirectly with the atomic sites through
275     the intermediary of the hull facets. Since each vertex (or atom)
276     provides one corner of a triangular facet, the force on the facets are
277     divided equally to each vertex. However, each vertex can participate
278     in multiple facets, so the resultant force is a sum over all facets
279     $f$ containing vertex $i$:
280 gezelter 3640 \begin{equation}
281     {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
282     } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
283     F}_f^{\mathrm ext}
284     \end{equation}
285    
286 gezelter 3652 The external pressure bath applies a force to the facets of the convex
287     hull in direct proportion to the area of the facet, while the thermal
288 gezelter 3660 coupling depends on the solvent temperature, viscosity and the size
289     and shape of each facet. The thermal interactions are expressed as a
290     standard Langevin description of the forces,
291 gezelter 3640 \begin{equation}
292     \begin{array}{rclclcl}
293     {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
294     & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
295     \end{array}
296     \end{equation}
297 gezelter 3665 Here, $A_f$ and $\hat{n}_f$ are the area and (outward-facing) normal
298     vectors for facet $f$, respectively. ${\mathbf v}_f(t)$ is the
299     velocity of the facet centroid,
300 gezelter 3652 \begin{equation}
301     {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
302     \end{equation}
303 gezelter 3660 and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
304     depends on the geometry and surface area of facet $f$ and the
305 gezelter 3689 viscosity of the bath. The resistance tensor is related to the
306 gezelter 3660 fluctuations of the random force, $\mathbf{R}(t)$, by the
307     fluctuation-dissipation theorem,
308 gezelter 3640 \begin{eqnarray}
309     \left< {\mathbf R}_f(t) \right> & = & 0 \\
310     \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
311 gezelter 3652 \Xi_f(t)\delta(t-t^\prime).
312     \label{eq:randomForce}
313 gezelter 3640 \end{eqnarray}
314    
315 gezelter 3665 Once the resistance tensor is known for a given facet, a stochastic
316 gezelter 3660 vector that has the properties in Eq. (\ref{eq:randomForce}) can be
317 gezelter 3665 calculated efficiently by carrying out a Cholesky decomposition to
318     obtain the square root matrix of the resistance tensor,
319 gezelter 3652 \begin{equation}
320     \Xi_f = {\bf S} {\bf S}^{T},
321     \label{eq:Cholesky}
322     \end{equation}
323     where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
324     vector with the statistics required for the random force can then be
325     obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
326     has elements chosen from a Gaussian distribution, such that:
327     \begin{equation}
328     \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
329     {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
330     \end{equation}
331     where $\delta t$ is the timestep in use during the simulation. The
332     random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
333     have the correct properties required by Eq. (\ref{eq:randomForce}).
334 gezelter 3640
335 gezelter 3689 Our treatment of the resistance tensor is approximate. $\Xi_f$ for a
336 gezelter 3660 rigid triangular plate would normally be treated as a $6 \times 6$
337 gezelter 3653 tensor that includes translational and rotational drag as well as
338 gezelter 3660 translational-rotational coupling. The computation of resistance
339 gezelter 3653 tensors for rigid bodies has been detailed
340 gezelter 3663 elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
341 gezelter 3653 but the standard approach involving bead approximations would be
342     prohibitively expensive if it were recomputed at each step in a
343     molecular dynamics simulation.
344    
345 gezelter 3665 Instead, we are utilizing an approximate resistance tensor obtained by
346     first constructing the Oseen tensor for the interaction of the
347     centroid of the facet ($f$) with each of the subfacets $\ell=1,2,3$,
348 gezelter 3653 \begin{equation}
349 gezelter 3665 T_{\ell f}=\frac{A_\ell}{8\pi\eta R_{\ell f}}\left(I +
350     \frac{\mathbf{R}_{\ell f}\mathbf{R}_{\ell f}^T}{R_{\ell f}^2}\right)
351 gezelter 3653 \end{equation}
352 gezelter 3665 Here, $A_\ell$ is the area of subfacet $\ell$ which is a triangle
353     containing two of the vertices of the facet along with the centroid.
354     $\mathbf{R}_{\ell f}$ is the vector between the centroid of facet $f$
355     and the centroid of sub-facet $\ell$, and $I$ is the ($3 \times 3$)
356     identity matrix. $\eta$ is the viscosity of the external bath.
357 gezelter 3653
358     \begin{figure}
359     \includegraphics[width=\linewidth]{hydro}
360 gezelter 3660 \caption{The resistance tensor $\Xi$ for a facet comprising sites $i$,
361     $j$, and $k$ is constructed using Oseen tensor contributions between
362     the centoid of the facet $f$ and each of the sub-facets ($i,f,j$),
363     ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets are
364     located at $1$, $2$, and $3$, and the area of each sub-facet is
365 gezelter 3653 easily computed using half the cross product of two of the edges.}
366     \label{hydro}
367     \end{figure}
368    
369 gezelter 3665 The tensors for each of the sub-facets are added together, and the
370     resulting matrix is inverted to give a $3 \times 3$ resistance tensor
371     for translations of the triangular facet,
372 gezelter 3653 \begin{equation}
373     \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
374     \end{equation}
375 gezelter 3667 Note that this treatment ignores rotations (and
376 gezelter 3660 translational-rotational coupling) of the facet. In compact systems,
377     the facets stay relatively fixed in orientation between
378     configurations, so this appears to be a reasonably good approximation.
379    
380 gezelter 3652 We have implemented this method by extending the Langevin dynamics
381 kstocke1 3716 integrator in our code, OpenMD.\cite{Meineke2005,open_md} At each
382 gezelter 3665 molecular dynamics time step, the following process is carried out:
383     \begin{enumerate}
384     \item The standard inter-atomic forces ($\nabla_iU$) are computed.
385 gezelter 3667 \item Delaunay triangulation is carried out using the current atomic
386 gezelter 3665 configuration.
387     \item The convex hull is computed and facets are identified.
388     \item For each facet:
389     \begin{itemize}
390 kstocke1 3690 \item[a.] The force from the pressure bath ($-\hat{n}_fPA_f$) is
391 gezelter 3665 computed.
392     \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
393     viscosity ($\eta$) of the bath.
394     \item[c.] Facet drag ($-\Xi_f(t) \mathbf{v}_f(t)$) forces are
395     computed.
396     \item[d.] Random forces ($\mathbf{R}_f(t)$) are computed using the
397     resistance tensor and the temperature ($T$) of the bath.
398     \end{itemize}
399     \item The facet forces are divided equally among the vertex atoms.
400     \item Atomic positions and velocities are propagated.
401     \end{enumerate}
402     The Delaunay triangulation and computation of the convex hull are done
403 kstocke1 3716 using calls to the qhull library.\cite{Q_hull} There is a minimal
404 gezelter 3665 penalty for computing the convex hull and resistance tensors at each
405     step in the molecular dynamics simulation (roughly 0.02 $\times$ cost
406     of a single force evaluation), and the convex hull is remarkably easy
407     to parallelize on distributed memory machines (see Appendix A).
408 gezelter 3652
409 gezelter 3640 \section{Tests \& Applications}
410 gezelter 3653 \label{sec:tests}
411 gezelter 3640
412 gezelter 3663 To test the new method, we have carried out simulations using the
413     Langevin Hull on: 1) a crystalline system (gold nanoparticles), 2) a
414 gezelter 3665 liquid droplet (SPC/E water),\cite{Berendsen1987} and 3) a
415 kstocke1 3695 heterogeneous mixture (gold nanoparticles in an SPC/E water droplet). In each case, we have computed properties that depend on the external applied pressure. Of particular interest for the single-phase systems is the isothermal compressibility,
416 gezelter 3660 \begin{equation}
417     \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
418     )_{T}.
419     \label{eq:BM}
420     \end{equation}
421    
422     One problem with eliminating periodic boundary conditions and
423     simulation boxes is that the volume of a three-dimensional point cloud
424 kstocke1 3695 is not well-defined. In order to compute the compressibility of a
425 gezelter 3660 bulk material, we make an assumption that the number density, $\rho =
426 kstocke1 3695 \frac{N}{V}$, is uniform within some region of the point cloud. The
427 gezelter 3660 compressibility can then be expressed in terms of the average number
428     of particles in that region,
429     \begin{equation}
430 gezelter 3665 \kappa_{T} = -\frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
431 kstocke1 3695 )_{T}.
432 gezelter 3660 \label{eq:BMN}
433     \end{equation}
434 kstocke1 3713 The region we used is a spherical volume of 20 \AA\ radius centered in
435 kstocke1 3716 the middle of the cluster with a roughly 25 \AA\ radius. $N$ is the average number of molecules
436 gezelter 3663 found within this region throughout a given simulation. The geometry
437 kstocke1 3716 of the region is arbitrary, and any bulk-like portion of the
438     cluster can be used to compute the compressibility.
439 gezelter 3660
440 gezelter 3665 One might assume that the volume of the convex hull could simply be
441     taken as the system volume $V$ in the compressibility expression
442     (Eq. \ref{eq:BM}), but this has implications at lower pressures (which
443     are explored in detail in the section on water droplets).
444 gezelter 3660
445 gezelter 3663 The metallic force field in use for the gold nanoparticles is the
446     quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} In all
447     simulations involving point charges, we utilized damped shifted-force
448     (DSF) electrostatics\cite{Fennell06} which is a variant of the Wolf
449     summation\cite{wolf:8254} that has been shown to provide good forces
450     and torques on molecular models for water in a computationally
451     efficient manner.\cite{Fennell06} The damping parameter ($\alpha$) was
452     set to 0.18 \AA$^{-1}$, and the cutoff radius was set to 12 \AA. The
453     Spohr potential was adopted in depicting the interaction between metal
454     atoms and the SPC/E water molecules.\cite{ISI:000167766600035}
455    
456 gezelter 3689 \subsection{Bulk Modulus of gold nanoparticles}
457 gezelter 3640
458 gezelter 3678 The compressibility (and its inverse, the bulk modulus) is well-known
459     for gold, and is captured well by the embedded atom method
460 gezelter 3689 (EAM)~\cite{PhysRevB.33.7983} potential and related multi-body force
461     fields. In particular, the quantum Sutton-Chen potential gets nearly
462     quantitative agreement with the experimental bulk modulus values, and
463     makes a good first test of how the Langevin Hull will perform at large
464     applied pressures.
465 gezelter 3663
466 gezelter 3678 The Sutton-Chen (SC) potentials are based on a model of a metal which
467     treats the nuclei and core electrons as pseudo-atoms embedded in the
468     electron density due to the valence electrons on all of the other
469 gezelter 3689 atoms in the system.\cite{Chen90} The SC potential has a simple form
470     that closely resembles the Lennard Jones potential,
471 gezelter 3678 \begin{equation}
472     \label{eq:SCP1}
473     U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
474     \end{equation}
475     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
476     \begin{equation}
477     \label{eq:SCP2}
478     V^{pair}_{ij}(r)=\left( \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}.
479     \end{equation}
480     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
481     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
482     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
483     the interactions between the valence electrons and the cores of the
484 gezelter 3722 pseudo-atoms. $D_{ij}$ and $D_{ii}$ set the appropriate overall energy
485 gezelter 3678 scale, $c_i$ scales the attractive portion of the potential relative
486     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
487     that assures a dimensionless form for $\rho$. These parameters are
488     tuned to various experimental properties such as the density, cohesive
489     energy, and elastic moduli for FCC transition metals. The quantum
490     Sutton-Chen (QSC) formulation matches these properties while including
491     zero-point quantum corrections for different transition
492 kstocke1 3716 metals.\cite{PhysRevB.59.3527,QSC2}
493 gezelter 3678
494     In bulk gold, the experimentally-measured value for the bulk modulus
495     is 180.32 GPa, while previous calculations on the QSC potential in
496 gezelter 3689 periodic-boundary simulations of the bulk crystal have yielded values
497 gezelter 3722 of 175.53 GPa.\cite{QSC2} Using the same force field, we have
498     performed a series of 1 ns simulations on gold nanoparticles of three
499 gezelter 3723 different radii: 20 \AA~ (1985 atoms), 30 \AA~ (6699 atoms), and 40
500     \AA~ (15707 atoms) utilizing the Langevin Hull at a variety of applied
501 gezelter 3722 pressures ranging from 0 -- 10 GPa. For the 40 \AA~ radius
502     nanoparticle we obtain a value of 177.55 GPa for the bulk modulus of
503     gold, in close agreement with both previous simulations and the
504     experimental bulk modulus reported for gold single
505     crystals.\cite{Collard1991} The smaller gold nanoparticles (30 and 20
506     \AA~ radii) have calculated bulk moduli of 215.58 and 208.86 GPa,
507     respectively, indicating that smaller nanoparticles are somewhat
508     stiffer (less compressible) than the larger nanoparticles. This
509     stiffening of the small nanoparticles may be related to their high
510     degree of surface curvature, resulting in a lower coordination number
511     of surface atoms relative to the the surface atoms in the 40 \AA~
512     radius particle.
513 gezelter 3678
514 gezelter 3722 We obtain a gold lattice constant of 4.051 \AA~ using the Langevin
515     Hull at 1 atm, close to the experimentally-determined value for bulk
516     gold and the value for gold simulated using the QSC potential and
517     periodic boundary conditions (4.079 \AA~ and 4.088\AA~,
518     respectively).\cite{QSC2} The slightly smaller calculated lattice
519     constant is most likely due to the presence of surface tension in the
520     non-periodic Langevin Hull cluster, an effect absent from a bulk
521     simulation. The specific heat of a 40 \AA~ gold nanoparticle under the
522     Langevin Hull at 1 atm is 24.914 $\mathrm {\frac{J}{mol \, K}}$, which
523     compares very well with the experimental value of 25.42 $\mathrm
524     {\frac{J}{mol \, K}}$.
525 kstocke1 3716
526 gezelter 3640 \begin{figure}
527 gezelter 3678 \includegraphics[width=\linewidth]{stacked}
528     \caption{The response of the internal pressure and temperature of gold
529     nanoparticles when first placed in the Langevin Hull
530     ($T_\mathrm{bath}$ = 300K, $P_\mathrm{bath}$ = 4 GPa), starting
531 gezelter 3665 from initial conditions that were far from the bath pressure and
532 gezelter 3678 temperature. The pressure response is rapid (after the breathing mode oscillations in the nanoparticle die out), and the rate of thermal equilibration depends on both exposed surface area (top panel) and the viscosity of the bath (middle panel).}
533 gezelter 3689 \label{fig:pressureResponse}
534 gezelter 3640 \end{figure}
535    
536 gezelter 3689 We note that the Langevin Hull produces rapidly-converging behavior
537     for structures that are started far from equilibrium. In
538     Fig. \ref{fig:pressureResponse} we show how the pressure and
539     temperature respond to the Langevin Hull for nanoparticles that were
540     initialized far from the target pressure and temperature. As
541     expected, the rate at which thermal equilibrium is achieved depends on
542 kstocke1 3713 the total surface area of the cluster exposed to the bath as well as
543 gezelter 3689 the bath viscosity. Pressure that is applied suddenly to a cluster
544     can excite breathing vibrations, but these rapidly damp out (on time
545 kstocke1 3695 scales of 30 -- 50 ps).
546 gezelter 3640
547     \subsection{Compressibility of SPC/E water clusters}
548    
549 gezelter 3660 Prior molecular dynamics simulations on SPC/E water (both in
550     NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
551     ensembles) have yielded values for the isothermal compressibility that
552     agree well with experiment.\cite{Fine1973} The results of two
553     different approaches for computing the isothermal compressibility from
554 kstocke1 3716 Langevin Hull simulations for pressures between 1 and 3000 atm are
555 gezelter 3660 shown in Fig. \ref{fig:compWater} along with compressibility values
556     obtained from both other SPC/E simulations and experiment.
557 kstocke1 3649
558 gezelter 3640 \begin{figure}
559 gezelter 3659 \includegraphics[width=\linewidth]{new_isothermalN}
560 kstocke1 3649 \caption{Compressibility of SPC/E water}
561 gezelter 3660 \label{fig:compWater}
562 gezelter 3640 \end{figure}
563    
564 gezelter 3660 Isothermal compressibility values calculated using the number density
565     (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
566 gezelter 3689 and previous simulation work throughout the 1 -- 1000 atm pressure
567 gezelter 3660 regime. Compressibilities computed using the Hull volume, however,
568     deviate dramatically from the experimental values at low applied
569 kstocke1 3715 pressures. The reason for this deviation is quite simple: at low
570 gezelter 3660 applied pressures, the liquid is in equilibrium with a vapor phase,
571     and it is entirely possible for one (or a few) molecules to drift away
572     from the liquid cluster (see Fig. \ref{fig:coneOfShame}). At low
573     pressures, the restoring forces on the facets are very gentle, and
574     this means that the hulls often take on relatively distorted
575     geometries which include large volumes of empty space.
576 kstocke1 3649
577 gezelter 3660 \begin{figure}
578 gezelter 3688 \includegraphics[width=\linewidth]{coneOfShame}
579 gezelter 3660 \caption{At low pressures, the liquid is in equilibrium with the vapor
580     phase, and isolated molecules can detach from the liquid droplet.
581 gezelter 3665 This is expected behavior, but the volume of the convex hull
582 gezelter 3689 includes large regions of empty space. For this reason,
583 gezelter 3662 compressibilities are computed using local number densities rather
584     than hull volumes.}
585 gezelter 3660 \label{fig:coneOfShame}
586     \end{figure}
587 kstocke1 3649
588 gezelter 3665 At higher pressures, the equilibrium strongly favors the liquid phase,
589     and the hull geometries are much more compact. Because of the
590     liquid-vapor effect on the convex hull, the regional number density
591     approach (Eq. \ref{eq:BMN}) provides more reliable estimates of the
592 gezelter 3667 compressibility.
593 kstocke1 3649
594 gezelter 3665 In both the traditional compressibility formula (Eq. \ref{eq:BM}) and
595     the number density version (Eq. \ref{eq:BMN}), multiple simulations at
596     different pressures must be done to compute the first derivatives. It
597     is also possible to compute the compressibility using the fluctuation
598     dissipation theorem using either fluctuations in the
599 kstocke1 3715 volume,\cite{Debenedetti1986}
600 kstocke1 3649 \begin{equation}
601 gezelter 3665 \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle
602     V \right \rangle ^{2}}{V \, k_{B} \, T},
603 gezelter 3689 \label{eq:BMVfluct}
604 kstocke1 3649 \end{equation}
605 gezelter 3665 or, equivalently, fluctuations in the number of molecules within the
606     fixed region,
607     \begin{equation}
608     \kappa_{T} = \frac{\left \langle N^{2} \right \rangle - \left \langle
609 kstocke1 3695 N \right \rangle ^{2}}{N \, k_{B} \, T}.
610 gezelter 3689 \label{eq:BMNfluct}
611 gezelter 3665 \end{equation}
612     Thus, the compressibility of each simulation can be calculated
613 gezelter 3689 entirely independently from other trajectories. Compressibility
614     calculations that rely on the hull volume will still suffer the
615     effects of the empty space due to the vapor phase; for this reason, we
616     recommend using the number density (Eq. \ref{eq:BMN}) or number
617     density fluctuations (Eq. \ref{eq:BMNfluct}) for computing
618 gezelter 3723 compressibilities. We obtained the results in
619     Fig. \ref{fig:compWater} using a sampling radius that was
620     approximately 80\% of the mean distance between the center of mass of
621     the cluster and the hull atoms. This ratio of sampling radius to
622     average hull radius excludes the problematic vapor phase on the
623     outside of the cluster while including enough of the liquid phase to
624     avoid poor statistics due to fluctuating local densities.
625 kstocke1 3649
626 gezelter 3722 A comparison of the oxygen-oxygen radial distribution functions for
627 gezelter 3723 SPC/E water simulated using both the Langevin Hull and more
628     traditional periodic boundary methods -- both at 1 atm and 300K --
629     reveals an understructuring of water in the Langevin Hull that
630     manifests as a slight broadening of the solvation shells. This effect
631     may be due to the introduction of a liquid-vapor interface in the
632     Langevin Hull simulations (an interface which is missing in most
633     periodic simulations of bulk water). Vapor-phase molecules contribute
634     a small but nearly flat portion of the radial distribution function.
635 kstocke1 3716
636 kstocke1 3649 \subsection{Molecular orientation distribution at cluster boundary}
637    
638 gezelter 3689 In order for a non-periodic boundary method to be widely applicable,
639 kstocke1 3690 it must be constructed in such a way that they allow a finite system
640 gezelter 3689 to replicate the properties of the bulk. Early non-periodic simulation
641     methods (e.g. hydrophobic boundary potentials) induced spurious
642     orientational correlations deep within the simulated
643 gezelter 3667 system.\cite{Lee1984,Belch1985} This behavior spawned many methods for
644 kstocke1 3716 fixing and characterizing the effects of artificial boundaries
645 gezelter 3667 including methods which fix the orientations of a set of edge
646     molecules.\cite{Warshel1978,King1989}
647 kstocke1 3649
648 gezelter 3667 As described above, the Langevin Hull does not require that the
649     orientation of molecules be fixed, nor does it utilize an explicitly
650 gezelter 3689 hydrophobic boundary, or orientational or radial constraints.
651     Therefore, the orientational correlations of the molecules in water
652     clusters are of particular interest in testing this method. Ideally,
653 kstocke1 3713 the water molecules on the surfaces of the clusters will have enough
654 gezelter 3689 mobility into and out of the center of the cluster to maintain
655 gezelter 3667 bulk-like orientational distribution in the absence of orientational
656     and radial constraints. However, since the number of hydrogen bonding
657     partners available to molecules on the exterior are limited, it is
658 gezelter 3689 likely that there will be an effective hydrophobicity of the hull.
659 kstocke1 3649
660 gezelter 3689 To determine the extent of these effects, we examined the
661 kstocke1 3690 orientations exhibited by SPC/E water in a cluster of 1372
662 gezelter 3689 molecules at 300 K and at pressures ranging from 1 -- 1000 atm. The
663 kstocke1 3690 orientational angle of a water molecule is described by
664 kstocke1 3649 \begin{equation}
665 gezelter 3640 \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
666     \end{equation}
667 gezelter 3667 where $\vec{r}_{i}$ is the vector between molecule {\it i}'s center of
668 gezelter 3689 mass and the cluster center of mass, and $\vec{\mu}_{i}$ is the vector
669     bisecting the H-O-H angle of molecule {\it i}. Bulk-like
670     distributions will result in $\langle \cos \theta \rangle$ values
671     close to zero. If the hull exhibits an overabundance of
672     externally-oriented oxygen sites, the average orientation will be
673     negative, while dangling hydrogen sites will result in positive
674     average orientations.
675 kstocke1 3649
676 gezelter 3667 Fig. \ref{fig:pAngle} shows the distribution of $\cos{\theta}$ values
677     for molecules in the interior of the cluster (squares) and for
678     molecules included in the convex hull (circles).
679 kstocke1 3649 \begin{figure}
680 gezelter 3640 \includegraphics[width=\linewidth]{pAngle}
681 gezelter 3667 \caption{Distribution of $\cos{\theta}$ values for molecules on the
682     interior of the cluster (squares) and for those participating in the
683 kstocke1 3694 convex hull (circles) at a variety of pressures. The Langevin Hull
684 gezelter 3667 exhibits minor dewetting behavior with exposed oxygen sites on the
685     hull water molecules. The orientational preference for exposed
686     oxygen appears to be independent of applied pressure. }
687     \label{fig:pAngle}
688 gezelter 3640 \end{figure}
689    
690 gezelter 3667 As expected, interior molecules (those not included in the convex
691     hull) maintain a bulk-like structure with a uniform distribution of
692     orientations. Molecules included in the convex hull show a slight
693     preference for values of $\cos{\theta} < 0.$ These values correspond
694     to molecules with oxygen directed toward the exterior of the cluster,
695 gezelter 3704 forming dangling hydrogen bond acceptor sites.
696 gezelter 3640
697 gezelter 3704 The orientational preference exhibited by water molecules on the hull
698     is significantly weaker than the preference caused by an explicit
699     hydrophobic bounding potential. Additionally, the Langevin Hull does
700     not require that the orientation of any molecules be fixed in order to
701     maintain bulk-like structure, even near the cluster surface.
702 kstocke1 3695
703 gezelter 3704 Previous molecular dynamics simulations of SPC/E liquid / vapor
704     interfaces using periodic boundary conditions have shown that
705     molecules on the liquid side of interface favor a similar orientation
706     where oxygen is directed away from the bulk.\cite{Taylor1996} These
707     simulations had well-defined liquid and vapor phase regions
708     equilibrium and it was observed that {\it vapor} molecules generally
709     had one hydrogen protruding from the surface, forming a dangling
710     hydrogen bond donor. Our water clusters do not have a true vapor
711     region, but rather a few transient molecules that leave the liquid
712     droplet (and which return to the droplet relatively quickly).
713     Although we cannot obtain an orientational preference of vapor phase
714     molecules in a Langevin Hull simulation, but we do agree with previous
715     estimates of the orientation of {\it liquid phase} molecules at the
716     interface.
717 kstocke1 3649
718 gezelter 3640 \subsection{Heterogeneous nanoparticle / water mixtures}
719    
720 kstocke1 3716 To further test the method, we simulated gold nanoparticles ($r = 18$
721 gezelter 3723 \AA~, 1433 atoms) solvated by explicit SPC/E water clusters (5000
722     molecules) using a model for the gold / water interactions that has
723     been used by Dou {\it et. al.} for investigating the separation of
724     water films near hot metal surfaces.\cite{ISI:000167766600035} The
725     Langevin Hull was used to sample pressures of 1, 2, 5, 10, 20, 50, 100
726     and 200 atm, while all simulations were done at a temperature of 300
727     K. At these temperatures and pressures, there is no observed
728     separation of the water film from the surface.
729 gezelter 3689
730 gezelter 3704 In Fig. \ref{fig:RhoR} we show the density of water and gold as a
731     function of the distance from the center of the nanoparticle. Higher
732     applied pressures appear to destroy structural correlations in the
733     outermost monolayer of the gold nanoparticle as well as in the water
734     at the near the metal / water interface. Simulations at increased
735     pressures exhibit significant overlap of the gold and water densities,
736     indicating a less well-defined interfacial surface.
737 kstocke1 3701
738 gezelter 3689 \begin{figure}
739 kstocke1 3699 \includegraphics[width=\linewidth]{RhoR}
740 gezelter 3704 \caption{Density profiles of gold and water at the nanoparticle
741     surface. Each curve has been normalized by the average density in
742 gezelter 3722 the bulk-like region available to the corresponding material.
743     Higher applied pressures de-structure both the gold nanoparticle
744     surface and water at the metal/water interface.}
745 gezelter 3689 \label{fig:RhoR}
746     \end{figure}
747    
748 gezelter 3704 At even higher pressures (500 atm and above), problems with the metal
749     - water interaction potential became quite clear. The model we are
750     using appears to have been parameterized for relatively low pressures;
751     it utilizes both shifted Morse and repulsive Morse potentials to model
752     the Au/O and Au/H interactions, respectively. The repulsive wall of
753     the Morse potential does not diverge quickly enough at short distances
754     to prevent water from diffusing into the center of the gold
755     nanoparticles. This behavior is likely not a realistic description of
756     the real physics of the situation. A better model of the gold-water
757 gezelter 3722 adsorption behavior would require harder repulsive walls to prevent
758     this behavior.
759 gezelter 3689
760 gezelter 3665 \section{Discussion}
761     \label{sec:discussion}
762 gezelter 3640
763 gezelter 3667 The Langevin Hull samples the isobaric-isothermal ensemble for
764 gezelter 3689 non-periodic systems by coupling the system to a bath characterized by
765     pressure, temperature, and solvent viscosity. This enables the
766 kstocke1 3690 simulation of heterogeneous systems composed of materials with
767 gezelter 3689 significantly different compressibilities. Because the boundary is
768     dynamically determined during the simulation and the molecules
769 kstocke1 3690 interacting with the boundary can change, the method inflicts minimal
770 gezelter 3689 perturbations on the behavior of molecules at the edges of the
771     simulation. Further work on this method will involve implicit
772     electrostatics at the boundary (which is missing in the current
773     implementation) as well as more sophisticated treatments of the
774     surface geometry (alpha
775 gezelter 3667 shapes\cite{EDELSBRUNNER:1994oq,EDELSBRUNNER:1995cj} and Tight
776     Cocone\cite{Dey:2003ts}). The non-convex hull geometries are
777     significantly more expensive ($\mathcal{O}(N^2)$) than the convex hull
778     ($\mathcal{O}(N \log N)$), but would enable the use of hull volumes
779     directly in computing the compressibility of the sample.
780    
781 gezelter 3663 \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
782 gezelter 3640
783 gezelter 3666 In order to use the Langevin Hull for simulations on parallel
784     computers, one of the more difficult tasks is to compute the bounding
785 gezelter 3689 surface, facets, and resistance tensors when the individual processors
786     have incomplete information about the entire system's topology. Most
787 gezelter 3666 parallel decomposition methods assign primary responsibility for the
788     motion of an atomic site to a single processor, and we can exploit
789     this to efficiently compute the convex hull for the entire system.
790    
791 gezelter 3667 The basic idea involves splitting the point cloud into
792     spatially-overlapping subsets and computing the convex hulls for each
793     of the subsets. The points on the convex hull of the entire system
794     are all present on at least one of the subset hulls. The algorithm
795     works as follows:
796 gezelter 3666 \begin{enumerate}
797     \item Each processor computes the convex hull for its own atomic sites
798 gezelter 3668 (left panel in Fig. \ref{fig:parallel}).
799 gezelter 3684 \item The Hull vertices from each processor are communicated to all of
800 gezelter 3666 the processors, and each processor assembles a complete list of hull
801     sites (this is much smaller than the original number of points in
802     the point cloud).
803 gezelter 3668 \item Each processor computes the global convex hull (right panel in
804 gezelter 3667 Fig. \ref{fig:parallel}) using only those points that are the union
805     of sites gathered from all of the subset hulls. Delaunay
806     triangulation is then done to obtain the facets of the global hull.
807 gezelter 3666 \end{enumerate}
808    
809     \begin{figure}
810 gezelter 3668 \includegraphics[width=\linewidth]{parallel}
811 gezelter 3666 \caption{When the sites are distributed among many nodes for parallel
812     computation, the processors first compute the convex hulls for their
813 gezelter 3668 own sites (dashed lines in left panel). The positions of the sites
814 gezelter 3669 that make up the subset hulls are then communicated to all
815 gezelter 3684 processors (middle panel). The convex hull of the system (solid line in
816     right panel) is the convex hull of the points on the union of the subset
817     hulls.}
818 gezelter 3668 \label{fig:parallel}
819 gezelter 3666 \end{figure}
820    
821     The individual hull operations scale with
822 gezelter 3667 $\mathcal{O}(\frac{n}{p}\log\frac{n}{p})$ where $n$ is the total
823     number of sites, and $p$ is the number of processors. These local
824 kstocke1 3690 hull operations create a set of $p$ hulls, each with approximately
825     $\frac{n}{3pr}$ sites for a cluster of radius $r$. The worst-case
826 gezelter 3667 communication cost for using a ``gather'' operation to distribute this
827     information to all processors is $\mathcal{O}( \alpha (p-1) + \frac{n
828     \beta (p-1)}{3 r p^2})$, while the final computation of the system
829     hull scales as $\mathcal{O}(\frac{n}{3r}\log\frac{n}{3r})$.
830 gezelter 3666
831 gezelter 3667 For a large number of atoms on a moderately parallel machine, the
832     total costs are dominated by the computations of the individual hulls,
833 kstocke1 3694 and communication of these hulls to create the Langevin Hull sees roughly
834 gezelter 3667 linear speed-up with increasing processor counts.
835    
836 gezelter 3663 \section*{Acknowledgments}
837 gezelter 3640 Support for this project was provided by the
838     National Science Foundation under grant CHE-0848243. Computational
839     time was provided by the Center for Research Computing (CRC) at the
840     University of Notre Dame.
841    
842 gezelter 3685 Molecular graphics images were produced using the UCSF Chimera package from
843     the Resource for Biocomputing, Visualization, and Informatics at the
844     University of California, San Francisco (supported by NIH P41 RR001081).
845 gezelter 3640 \newpage
846    
847     \bibliography{langevinHull}
848    
849     \end{doublespace}
850     \end{document}