ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
Revision: 3662
Committed: Thu Oct 21 15:43:22 2010 UTC (13 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 29474 byte(s)
Log Message:
edits

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. As long as the material in
84 the simulation box is essentially a bulk-like liquid which has a
85 relatively uniform compressibility, the standard affine transform
86 approach provides an excellent way of adjusting the volume of the
87 system and applying pressure directly via the interactions between
88 atomic sites.
89
90 The problem with this approach becomes apparent when the material
91 being simulated is an inhomogeneous mixture in which portions of the
92 simulation box are incompressible relative to other portions.
93 Examples include simulations of metallic nanoparticles in liquid
94 environments, proteins at interfaces, as well as other multi-phase or
95 interfacial environments. In these cases, the affine transform of
96 atomic coordinates will either cause numerical instability when the
97 sites in the incompressible medium collide with each other, or lead to
98 inefficient sampling of system volumes if the barostat is set slow
99 enough to avoid the instabilities in the incompressible region.
100
101 \begin{figure}
102 \includegraphics[width=\linewidth]{AffineScale2}
103 \caption{Affine Scaling constant pressure methods use box-length
104 scaling to adjust the volume to adjust to under- or over-pressure
105 conditions. In a system with a uniform compressibility (e.g. bulk
106 fluids) these methods can work well. In systems containing
107 heterogeneous mixtures, the affine scaling moves required to adjust
108 the pressure in the high-compressibility regions can cause molecules
109 in low compressibility regions to collide.}
110 \label{affineScale}
111 \end{figure}
112
113 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 volume either requires effective solute concentrations that are much
117 higher than desirable, or unreasonable system sizes to avoid this
118 effect. For example, calculations using typical hydration shells
119 solvating a protein under periodic boundary conditions are quite
120 expensive. [CALCULATE EFFECTIVE PROTEIN CONCENTRATIONS IN TYPICAL
121 SIMULATIONS]
122
123 There have been a number of other approaches to explicit
124 non-periodicity that focus on constant or nearly-constant {\it volume}
125 conditions while maintaining bulk-like behavior. Berkowitz and
126 McCammon introduced a stochastic (Langevin) boundary layer inside a
127 region of fixed molecules which effectively enforces constant
128 temperature and volume (NVT) conditions.\cite{Berkowitz1982} In this
129 approach, the stochastic and fixed regions were defined relative to a
130 central atom. Brooks and Karplus extended this method to include
131 deformable stochastic boundaries.\cite{iii:6312} The stochastic
132 boundary approach has been used widely for protein
133 simulations. [CITATIONS NEEDED]
134
135 The electrostatic and dispersive behavior near the boundary has long
136 been a cause for concern. King and Warshel introduced a surface
137 constrained all-atom solvent (SCAAS) which included polarization
138 effects of a fixed spherical boundary to mimic bulk-like behavior
139 without periodic boundaries.\cite{king:3647} In the SCAAS model, a
140 layer of fixed solvent molecules surrounds the solute and any explicit
141 solvent, and this in turn is surrounded by a continuum dielectric.
142 MORE HERE. WHAT DID THEY FIND?
143
144 Beglov and Roux developed a boundary model in which the hard sphere
145 boundary has a radius that varies with the instantaneous configuration
146 of the solute (and solvent) molecules.\cite{beglov:9050} This model
147 contains a clear pressure and surface tension contribution to the free
148 energy which XXX.
149
150 Restraining {\it potentials} introduce repulsive potentials at the
151 surface of a sphere or other geometry. The solute and any explicit
152 solvent are therefore restrained inside this potential. Often the
153 potentials include a weak short-range attraction to maintain the
154 correct density at the boundary. Beglov and Roux have also introduced
155 a restraining boundary potential which relaxes dynamically depending
156 on the solute geometry and the force the explicit system exerts on the
157 shell.\cite{Beglov:1995fk}
158
159 Recently, Krilov {\it et al.} introduced a flexible boundary model
160 that uses a Lennard-Jones potential between the solvent molecules and
161 a boundary which is determined dynamically from the position of the
162 nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This approach allows
163 the confining potential to prevent solvent molecules from migrating
164 too far from the solute surface, while providing a weak attractive
165 force pulling the solvent molecules towards a fictitious bulk solvent.
166 Although this approach is appealing and has physical motivation,
167 nanoparticles do not deform far from their original geometries even at
168 temperatures which vaporize the nearby solvent. For the systems like
169 the one described, the flexible boundary model will be nearly
170 identical to a fixed-volume restraining potential.
171
172 The approach of Kohanoff, Caro, and Finnis is the most promising of
173 the methods for introducing both constant pressure and temperature
174 into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
175 This method is based on standard Langevin dynamics, but the Brownian
176 or random forces are allowed to act only on peripheral atoms and exert
177 force in a direction that is inward-facing relative to the facets of a
178 closed bounding surface. The statistical distribution of the random
179 forces are uniquely tied to the pressure in the external reservoir, so
180 the method can be shown to sample the isobaric-isothermal ensemble.
181 Kohanoff {\it et al.} used a Delaunay tessellation to generate a
182 bounding surface surrounding the outermost atoms in the simulated
183 system. This is not the only possible triangulated outer surface, but
184 guarantees that all of the random forces point inward towards the
185 cluster.
186
187 In the following sections, we extend and generalize the approach of
188 Kohanoff, Caro, and Finnis. The new method, which we are calling the
189 ``Langevin Hull'' applies the external pressure, Langevin drag, and
190 random forces on the facets of the {\it hull itself} instead of the
191 atomic sites comprising the vertices of the hull. This allows us to
192 decouple the external pressure contribution from the drag and random
193 force. Section \ref{sec:meth}
194
195 \section{Methodology}
196 \label{sec:meth}
197
198 We have developed a new method which uses an external bath at a fixed
199 constant pressure ($P$) and temperature ($T$). This bath interacts
200 only with the objects on the exterior hull of the system. Defining
201 the hull of the simulation is done in a manner similar to the approach
202 of Kohanoff, Caro and Finnis.\cite{Kohanoff:2005qm} That is, any
203 instantaneous configuration of the atoms in the system is considered
204 as a point cloud in three dimensional space. Delaunay triangulation
205 is used to find all facets between coplanar neighbors.\cite{DELAUNAY}
206 In highly symmetric point clouds, facets can contain many atoms, but
207 in all but the most symmetric of cases the facets are simple triangles
208 in 3-space that contain exactly three atoms.
209
210 The convex hull is the set of facets that have {\it no concave
211 corners} at an atomic site. This eliminates all facets on the
212 interior of the point cloud, leaving only those exposed to the
213 bath. Sites on the convex hull are dynamic. As molecules re-enter the
214 cluster, all interactions between atoms on that molecule and the
215 external bath are removed. Since the edge is determined dynamically
216 as the simulation progresses, no {\it a priori} geometry is
217 defined. The pressure and temperature bath interacts {\it directly}
218 with the atoms on the edge and not with atoms interior to the
219 simulation.
220
221
222 \begin{figure}
223 \includegraphics[width=\linewidth]{hullSample}
224 \caption{The external temperature and pressure bath interacts only
225 with those atoms on the convex hull (grey surface). The hull is
226 computed dynamically at each time step, and molecules dynamically
227 move between the interior (Newtonian) region and the Langevin hull.}
228 \label{fig:hullSample}
229 \end{figure}
230
231
232 Atomic sites in the interior of the point cloud move under standard
233 Newtonian dynamics,
234 \begin{equation}
235 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
236 \label{eq:Newton}
237 \end{equation}
238 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
239 instantaneous velocity of site $i$ at time $t$, and $U$ is the total
240 potential energy. For atoms on the exterior of the cluster
241 (i.e. those that occupy one of the vertices of the convex hull), the
242 equation of motion is modified with an external force, ${\mathbf
243 F}_i^{\mathrm ext}$,
244 \begin{equation}
245 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
246 \end{equation}
247
248 The external bath interacts directly with the facets of the convex
249 hull. Since each vertex (or atom) provides one corner of a triangular
250 facet, the force on the facets are divided equally to each vertex.
251 However, each vertex can participate in multiple facets, so the resultant
252 force is a sum over all facets $f$ containing vertex $i$:
253 \begin{equation}
254 {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
255 } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
256 F}_f^{\mathrm ext}
257 \end{equation}
258
259 The external pressure bath applies a force to the facets of the convex
260 hull in direct proportion to the area of the facet, while the thermal
261 coupling depends on the solvent temperature, viscosity and the size
262 and shape of each facet. The thermal interactions are expressed as a
263 standard Langevin description of the forces,
264 \begin{equation}
265 \begin{array}{rclclcl}
266 {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
267 & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
268 \end{array}
269 \end{equation}
270 Here, $A_f$ and $\hat{n}_f$ are the area and normal vectors for facet
271 $f$, respectively. ${\mathbf v}_f(t)$ is the velocity of the facet
272 centroid,
273 \begin{equation}
274 {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
275 \end{equation}
276 and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
277 depends on the geometry and surface area of facet $f$ and the
278 viscosity of the fluid. The resistance tensor is related to the
279 fluctuations of the random force, $\mathbf{R}(t)$, by the
280 fluctuation-dissipation theorem,
281 \begin{eqnarray}
282 \left< {\mathbf R}_f(t) \right> & = & 0 \\
283 \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
284 \Xi_f(t)\delta(t-t^\prime).
285 \label{eq:randomForce}
286 \end{eqnarray}
287
288 Once the resistance tensor is known for a given facet a stochastic
289 vector that has the properties in Eq. (\ref{eq:randomForce}) can be
290 done efficiently by carrying out a Cholesky decomposition to obtain
291 the square root matrix of the resistance tensor,
292 \begin{equation}
293 \Xi_f = {\bf S} {\bf S}^{T},
294 \label{eq:Cholesky}
295 \end{equation}
296 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
297 vector with the statistics required for the random force can then be
298 obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
299 has elements chosen from a Gaussian distribution, such that:
300 \begin{equation}
301 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
302 {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
303 \end{equation}
304 where $\delta t$ is the timestep in use during the simulation. The
305 random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
306 have the correct properties required by Eq. (\ref{eq:randomForce}).
307
308 Our treatment of the resistance tensor is approximate. $\Xi$ for a
309 rigid triangular plate would normally be treated as a $6 \times 6$
310 tensor that includes translational and rotational drag as well as
311 translational-rotational coupling. The computation of resistance
312 tensors for rigid bodies has been detailed
313 elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun2008}
314 but the standard approach involving bead approximations would be
315 prohibitively expensive if it were recomputed at each step in a
316 molecular dynamics simulation.
317
318 We are utilizing an approximate resistance tensor obtained by first
319 constructing the Oseen tensor for the interaction of the centroid of
320 the facet ($f$) with each of the subfacets $j$,
321 \begin{equation}
322 T_{jf}=\frac{A_j}{8\pi\eta R_{jf}}\left(I +
323 \frac{\mathbf{R}_{jf}\mathbf{R}_{jf}^T}{R_{jf}^2}\right)
324 \end{equation}
325 Here, $A_j$ is the area of subfacet $j$ which is a triangle containing
326 two of the vertices of the facet along with the centroid.
327 $\mathbf{R}_{jf}$ is the vector between the centroid of facet $f$ and
328 the centroid of sub-facet $j$, and $I$ is the ($3 \times 3$) identity
329 matrix. $\eta$ is the viscosity of the external bath.
330
331 \begin{figure}
332 \includegraphics[width=\linewidth]{hydro}
333 \caption{The resistance tensor $\Xi$ for a facet comprising sites $i$,
334 $j$, and $k$ is constructed using Oseen tensor contributions between
335 the centoid of the facet $f$ and each of the sub-facets ($i,f,j$),
336 ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets are
337 located at $1$, $2$, and $3$, and the area of each sub-facet is
338 easily computed using half the cross product of two of the edges.}
339 \label{hydro}
340 \end{figure}
341
342 The Oseen tensors for each of the sub-facets are added together, and
343 the resulting matrix is inverted to give a $3 \times 3$ resistance
344 tensor for translations of the triangular facet,
345 \begin{equation}
346 \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
347 \end{equation}
348 Note that this treatment explicitly ignores rotations (and
349 translational-rotational coupling) of the facet. In compact systems,
350 the facets stay relatively fixed in orientation between
351 configurations, so this appears to be a reasonably good approximation.
352
353 We have implemented this method by extending the Langevin dynamics
354 integrator in our code, OpenMD.\cite{Meineke2005,openmd} The Delaunay
355 triangulation and computation of the convex hull are done using calls
356 to the qhull library.\cite{qhull} There is a moderate penalty for
357 computing the convex hull at each step in the molecular dynamics
358 simulation (HOW MUCH?), but the convex hull is remarkably easy to
359 parallelize on distributed memory machines (see Appendix A).
360
361 \section{Tests \& Applications}
362 \label{sec:tests}
363
364 In order to test this method, we have carried out simulations using
365 the Langevin Hull on a crystalline system (gold nanoparticles), a
366 liquid droplet (SPC/E water), and a heterogeneous mixture (gold
367 nanoparticles in a water droplet). In each case, we have computed
368 properties that depend on the external applied pressure. Of
369 particular interest for the single-phase systems is the bulk modulus,
370 \begin{equation}
371 \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
372 )_{T}.
373 \label{eq:BM}
374 \end{equation}
375
376 One problem with eliminating periodic boundary conditions and
377 simulation boxes is that the volume of a three-dimensional point cloud
378 is not well-defined. In order to compute the compressibility of a
379 bulk material, we make an assumption that the number density, $\rho =
380 \frac{N}{V}$, is uniform within some region of the cloud. The
381 compressibility can then be expressed in terms of the average number
382 of particles in that region,
383 \begin{equation}
384 \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
385 )_{T}
386 \label{eq:BMN}
387 \end{equation}
388 The region we pick is a spherical volume of 10 \AA radius centered in
389 the middle of the cluster. The geometry and size of the region is
390 arbitrary, and any bulk-like portion of the cluster can be used to
391 compute the bulk modulus.
392
393 One might also assume that the volume of the convex hull could be
394 taken as the system volume in the compressibility expression
395 (Eq. \ref{eq:BM}), but this has implications at lower pressures (which
396 are explored in detail in the section on water droplets).
397
398 \subsection{Bulk modulus of gold nanoparticles}
399
400 \begin{figure}
401 \includegraphics[width=\linewidth]{pressure_tb}
402 \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
403 pressure = 4 GPa}
404 \label{pressureResponse}
405 \end{figure}
406
407 \begin{figure}
408 \includegraphics[width=\linewidth]{temperature_tb}
409 \caption{Temperature equilibration depends on surface area and bath
410 viscosity. Target Temperature = 300K}
411 \label{temperatureResponse}
412 \end{figure}
413
414 \begin{equation}
415 \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
416 P}\right)
417 \end{equation}
418
419 \begin{figure}
420 \includegraphics[width=\linewidth]{compress_tb}
421 \caption{Isothermal Compressibility (18 \AA gold nanoparticle)}
422 \label{temperatureResponse}
423 \end{figure}
424
425 \subsection{Compressibility of SPC/E water clusters}
426
427 Prior molecular dynamics simulations on SPC/E water (both in
428 NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
429 ensembles) have yielded values for the isothermal compressibility that
430 agree well with experiment.\cite{Fine1973} The results of two
431 different approaches for computing the isothermal compressibility from
432 Langevin Hull simulations for pressures between 1 and 6500 atm are
433 shown in Fig. \ref{fig:compWater} along with compressibility values
434 obtained from both other SPC/E simulations and experiment.
435 Compressibility values from all references are for applied pressures
436 within the range 1 - 1000 atm.
437
438 \begin{figure}
439 \includegraphics[width=\linewidth]{new_isothermalN}
440 \caption{Compressibility of SPC/E water}
441 \label{fig:compWater}
442 \end{figure}
443
444 Isothermal compressibility values calculated using the number density
445 (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
446 and previous simulation work throughout the 1 - 1000 atm pressure
447 regime. Compressibilities computed using the Hull volume, however,
448 deviate dramatically from the experimental values at low applied
449 pressures. The reason for this deviation is quite simple; at low
450 applied pressures, the liquid is in equilibrium with a vapor phase,
451 and it is entirely possible for one (or a few) molecules to drift away
452 from the liquid cluster (see Fig. \ref{fig:coneOfShame}). At low
453 pressures, the restoring forces on the facets are very gentle, and
454 this means that the hulls often take on relatively distorted
455 geometries which include large volumes of empty space.
456
457 \begin{figure}
458 \includegraphics[width=\linewidth]{flytest2}
459 \caption{At low pressures, the liquid is in equilibrium with the vapor
460 phase, and isolated molecules can detach from the liquid droplet.
461 This is expected behavior, but the reported volume of the convex
462 hull includes large regions of empty space. For this reason,
463 compressibilities are computed using local number densities rather
464 than hull volumes.}
465 \label{fig:coneOfShame}
466 \end{figure}
467
468 At higher pressures, the equilibrium favors the liquid phase, and the
469 hull geometries are much more compact. Because of the liquid-vapor
470 effect on the convex hull, the regional number density approach
471 (Eq. \ref{eq:BMN}) provides more reliable estimates of the bulk
472 modulus.
473
474 We initially used the classic compressibility formula 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.
475
476 Regardless of the difficulty in obtaining accurate hull
477 volumes at low temperature and pressures, the Langevin Hull NPT method
478 provides reasonable isothermal compressibility values for water
479 through a large range of pressures.
480
481 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
482
483 \begin{equation}
484 \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle V \right \rangle ^{2}}{V \, k_{B} \, T}
485 \end{equation}
486
487 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.
488
489
490 \subsection{Molecular orientation distribution at cluster boundary}
491
492 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.
493
494 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.
495
496 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.
497
498 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.
499
500 The orientation of a water molecule is described by
501
502 \begin{equation}
503 \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
504 \end{equation}
505
506 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}.
507
508 \begin{figure}
509 \includegraphics[width=\linewidth]{g_r_theta}
510 \caption{Definition of coordinates}
511 \label{coords}
512 \end{figure}
513
514 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).
515
516 \begin{figure}
517 \includegraphics[width=\linewidth]{pAngle}
518 \caption{SPC/E water clusters: only minor dewetting at the boundary}
519 \label{pAngle}
520 \end{figure}
521
522 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.
523
524 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.
525
526 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.
527
528
529 \subsection{Heterogeneous nanoparticle / water mixtures}
530
531
532 \section{Appendix A: Hydrodynamic tensor for triangular facets}
533
534 \section{Appendix B: Computing Convex Hulls on Parallel Computers}
535
536 \section{Acknowledgments}
537 Support for this project was provided by the
538 National Science Foundation under grant CHE-0848243. Computational
539 time was provided by the Center for Research Computing (CRC) at the
540 University of Notre Dame.
541
542 \newpage
543
544 \bibliography{langevinHull}
545
546 \end{doublespace}
547 \end{document}