ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
Revision: 3653
Committed: Tue Oct 19 16:13:01 2010 UTC (13 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 27022 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 a constant pressure and
199 temperature bath. This bath interacts only with the objects that are
200 currently at the edge of the system. Since the edge is determined
201 dynamically as the simulation progresses, no {\it a priori} geometry
202 is defined. The pressure and temperature bath interacts {\it
203 directly} with the atoms on the edge and not with atoms interior to
204 the simulation. This means that there are no affine transforms
205 required. There are also no fictitious particles or bounding
206 potentials used in this approach.
207
208 The basics of the method are as follows. The simulation starts as a
209 collection of atomic locations in three dimensions (a point cloud).
210 Delaunay triangulation is used to find all facets between coplanar
211 neighbors. In highly symmetric point clouds, facets can contain many
212 atoms, but in all but the most symmetric of cases one might experience
213 in a molecular dynamics simulation, the facets are simple triangles in
214 3-space that contain exactly three atoms.
215
216 The convex hull is the set of facets that have {\it no concave
217 corners} at an atomic site. This eliminates all facets on the
218 interior of the point cloud, leaving only those exposed to the
219 bath. Sites on the convex hull are dynamic. As molecules re-enter the
220 cluster, all interactions between atoms on that molecule and the
221 external bath are removed.
222
223 For atomic sites in the interior of the point cloud, the equations of
224 motion are simple Newtonian dynamics,
225 \begin{equation}
226 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
227 \label{eq:Newton}
228 \end{equation}
229 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
230 instantaneous velocity of site $i$ at time $t$, and $U$ is the total
231 potential energy. For atoms on the exterior of the cluster
232 (i.e. those that occupy one of the vertices of the convex hull), the
233 equation of motion is modified with an external force, ${\mathbf
234 F}_i^{\mathrm ext}$,
235 \begin{equation}
236 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
237 \end{equation}
238
239 The external bath interacts directly with the facets of the convex
240 hull. Since each vertex (or atom) provides one corner of a triangular
241 facet, the force on the facets are divided equally to each vertex.
242 However, each vertex can participate in multiple facets, so the resultant
243 force is a sum over all facets $f$ containing vertex $i$:
244 \begin{equation}
245 {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
246 } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
247 F}_f^{\mathrm ext}
248 \end{equation}
249
250 The external pressure bath applies a force to the facets of the convex
251 hull in direct proportion to the area of the facet, while the thermal
252 coupling depends on the solvent temperature, friction and the size and
253 shape of each facet. The thermal interactions are expressed as a
254 typical Langevin description of the forces,
255 \begin{equation}
256 \begin{array}{rclclcl}
257 {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
258 & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
259 \end{array}
260 \end{equation}
261 Here, $P$ is the external pressure, $A_f$ and $\hat{n}_f$ are the area
262 and normal vectors for facet $f$, respectively. ${\mathbf v}_f(t)$ is
263 the velocity of the facet,
264 \begin{equation}
265 {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
266 \end{equation}
267 and $\Xi_f(t)$ is an approximate ($3 \times 3$) hydrodynamic tensor
268 that depends on the geometry and surface area of facet $f$ and the
269 viscosity of the fluid (See Appendix A). The hydrodynamic tensor is
270 related to the fluctuations of the random force, $\mathbf{R}(t)$, by
271 the fluctuation-dissipation theorem,
272 \begin{eqnarray}
273 \left< {\mathbf R}_f(t) \right> & = & 0 \\
274 \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
275 \Xi_f(t)\delta(t-t^\prime).
276 \label{eq:randomForce}
277 \end{eqnarray}
278
279 Once the hydrodynamic tensor is known for a given facet (see Appendix
280 A) obtaining a stochastic vector that has the properties in
281 Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
282 one-time Cholesky decomposition to obtain the square root matrix of
283 the resistance tensor,
284 \begin{equation}
285 \Xi_f = {\bf S} {\bf S}^{T},
286 \label{eq:Cholesky}
287 \end{equation}
288 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
289 vector with the statistics required for the random force can then be
290 obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
291 has elements chosen from a Gaussian distribution, such that:
292 \begin{equation}
293 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
294 {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
295 \end{equation}
296 where $\delta t$ is the timestep in use during the simulation. The
297 random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
298 have the correct properties required by Eq. (\ref{eq:randomForce}).
299
300 Our treatment of the hydrodynamic tensor must be approximate. $\Xi$
301 for a triangular plate would normally be treated as a $6 \times 6$
302 tensor that includes translational and rotational drag as well as
303 translational-rotational coupling. The computation of hydrodynamic
304 tensors for rigid bodies has been detailed
305 elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun2008}
306 but the standard approach involving bead approximations would be
307 prohibitively expensive if it were recomputed at each step in a
308 molecular dynamics simulation.
309
310 We are utilizing an approximate hydrodynamic tensor obtained by first
311 constructing the Oseen tensor for the interaction of the centroid of
312 the facet ($f$) with each of the subfacets $j$,
313 \begin{equation}
314 T_{jf}=\frac{A_j}{8\pi\eta R_{jf}}\left(I +
315 \frac{\mathbf{R}_{jf}\mathbf{R}_{jf}^T}{R_{jf}^2}\right)
316 \end{equation}
317 Here, $A_j$ is the area of subfacet $j$ which is a triangle containing
318 two of the vertices of the facet along with the centroid.
319 $\mathbf{R}_{jf}$ is the vector between the centroid of facet $f$ and
320 the centroid of sub-facet $j$, and $I$ is the ($3 \times 3$) identity
321 matrix. $\eta$ is the viscosity of the external bath.
322
323 \begin{figure}
324 \includegraphics[width=\linewidth]{hydro}
325 \caption{The hydrodynamic tensor $\Xi$ for a facet comprising sites $i$,
326 $j$, and $k$ is constructed using Oseen tensor contributions
327 between the centoid of the facet $f$ and each of the sub-facets
328 ($i,f,j$), ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets
329 are located at $1$, $2$, and $3$, and the area of each sub-facet is
330 easily computed using half the cross product of two of the edges.}
331 \label{hydro}
332 \end{figure}
333
334 The Oseen tensors for each of the sub-facets are summed, and the
335 resulting matrix is inverted to give a $3 \times 3$ hydrodynamic
336 tensor for translations of the triangular plate,
337 \begin{equation}
338 \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
339 \end{equation}
340 We have implemented this method by extending the Langevin dynamics
341 integrator in our group code, OpenMD.\cite{Meineke2005,openmd} There
342 is a moderate penalty for computing the convex hull at each step in
343 the molecular dynamics simulation (HOW MUCH?), but the convex hull is
344 remarkably easy to parallelize on distributed memory machines (see
345 Appendix B).
346
347 \section{Tests \& Applications}
348 \label{sec:tests}
349
350 \subsection{Bulk modulus of gold nanoparticles}
351
352 \begin{figure}
353 \includegraphics[width=\linewidth]{pressure_tb}
354 \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
355 pressure = 4 GPa}
356 \label{pressureResponse}
357 \end{figure}
358
359 \begin{figure}
360 \includegraphics[width=\linewidth]{temperature_tb}
361 \caption{Temperature equilibration depends on surface area and bath
362 viscosity. Target Temperature = 300K}
363 \label{temperatureResponse}
364 \end{figure}
365
366 \begin{equation}
367 \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
368 P}\right)
369 \end{equation}
370
371 \begin{figure}
372 \includegraphics[width=\linewidth]{compress_tb}
373 \caption{Isothermal Compressibility (18 \AA gold nanoparticle)}
374 \label{temperatureResponse}
375 \end{figure}
376
377 \subsection{Compressibility of SPC/E water clusters}
378
379 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.
380
381 \begin{figure}
382 \includegraphics[width=\linewidth]{new_isothermal}
383 \caption{Compressibility of SPC/E water}
384 \label{compWater}
385 \end{figure}
386
387 We initially used the classic compressibility formula
388
389 \begin{equation}
390 \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right )_{T}
391 \end{equation}
392
393 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.
394
395 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
396
397 \begin{equation}
398 \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle V \right \rangle ^{2}}{V \, k_{B} \, T}
399 \end{equation}
400
401 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.
402
403 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
404
405 \begin{equation}
406 \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right )_{T}
407 \end{equation}
408
409 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.
410
411 \subsection{Molecular orientation distribution at cluster boundary}
412
413 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.
414
415 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.
416
417 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.
418
419 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.
420
421 The orientation of a water molecule is described by
422
423 \begin{equation}
424 \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
425 \end{equation}
426
427 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}.
428
429 \begin{figure}
430 \includegraphics[width=\linewidth]{g_r_theta}
431 \caption{Definition of coordinates}
432 \label{coords}
433 \end{figure}
434
435 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).
436
437 \begin{figure}
438 \includegraphics[width=\linewidth]{pAngle}
439 \caption{SPC/E water clusters: only minor dewetting at the boundary}
440 \label{pAngle}
441 \end{figure}
442
443 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.
444
445 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.
446
447 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.
448
449
450 \subsection{Heterogeneous nanoparticle / water mixtures}
451
452
453 \section{Appendix A: Hydrodynamic tensor for triangular facets}
454
455 \section{Appendix B: Computing Convex Hulls on Parallel Computers}
456
457 \section{Acknowledgments}
458 Support for this project was provided by the
459 National Science Foundation under grant CHE-0848243. Computational
460 time was provided by the Center for Research Computing (CRC) at the
461 University of Notre Dame.
462
463 \newpage
464
465 \bibliography{langevinHull}
466
467 \end{doublespace}
468 \end{document}