ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
Revision: 3688
Committed: Thu Nov 11 17:13:49 2010 UTC (13 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 38334 byte(s)
Log Message:
new figures

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     \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     materials with different compressibilities. These are systems that
46     are problematic for traditional affine transform methods. The
47     Langevin Hull does not suffer from the edge effects of boundary
48     potential methods, and allows realistic treatment of both external
49     pressure and thermal conductivity due to the presence of an implicit
50     solvent. We apply this method to several different systems
51     including bare metal nanoparticles, nanoparticles in an explicit
52     solvent, as well as clusters of liquid water. The predicted
53     mechanical properties of these systems are in good agreement with
54     experimental data and 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 3684 expensive. A 62 $\AA^3$ box of water solvating a moderately small
121     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 gezelter 3684 Typically protein concentrations in the cell are on the order of
125     160-310 mg/ml,\cite{Brown1991195} and the factor of 20 difference
126     between simulations and the cellular environment may have significant
127     effects on the structure and dynamics of simulated protein structures.
128    
129    
130 gezelter 3665 \subsection*{Boundary Methods}
131 gezelter 3667 There have been a number of approaches to handle simulations of
132     explicitly non-periodic systems that focus on constant or
133     nearly-constant {\it volume} conditions while maintaining bulk-like
134     behavior. Berkowitz and McCammon introduced a stochastic (Langevin)
135     boundary layer inside a region of fixed molecules which effectively
136     enforces constant temperature and volume (NVT)
137     conditions.\cite{Berkowitz1982} In this approach, the stochastic and
138     fixed regions were defined relative to a central atom. Brooks and
139     Karplus extended this method to include deformable stochastic
140     boundaries.\cite{iii:6312} The stochastic boundary approach has been
141     used widely for protein simulations. [CITATIONS NEEDED]
142 gezelter 3640
143 gezelter 3653 The electrostatic and dispersive behavior near the boundary has long
144 gezelter 3665 been a cause for concern when performing simulations of explicitly
145     non-periodic systems. Early work led to the surface constrained soft
146     sphere dipole model (SCSSD)\cite{Warshel1978} in which the surface
147     molecules are fixed in a random orientation representative of the bulk
148     solvent structural properties. Belch {\it et al.}\cite{Belch1985}
149     simulated clusters of TIPS2 water surrounded by a hydrophobic bounding
150     potential. The spherical hydrophobic boundary induced dangling
151     hydrogen bonds at the surface that propagated deep into the cluster,
152     affecting most of molecules in the simulation. This result echoes an
153     earlier study which showed that an extended planar hydrophobic surface
154     caused orientational preference at the surface which extended
155     relatively deep (7 \r{A}) into the liquid simulation
156     cell.\cite{Lee1984} The surface constrained all-atom solvent (SCAAS)
157     model \cite{King1989} improved upon its SCSSD predecessor. The SCAAS
158     model utilizes a polarization constraint which is applied to the
159     surface molecules to maintain bulk-like structure at the cluster
160     surface. A radial constraint is used to maintain the desired bulk
161     density of the liquid. Both constraint forces are applied only to a
162     pre-determined number of the outermost molecules.
163 gezelter 3640
164 gezelter 3665 Beglov and Roux have developed a boundary model in which the hard
165     sphere boundary has a radius that varies with the instantaneous
166     configuration of the solute (and solvent) molecules.\cite{beglov:9050}
167     This model contains a clear pressure and surface tension contribution
168     to the free energy which XXX.
169 gezelter 3640
170 gezelter 3665 \subsection*{Restraining Potentials}
171 gezelter 3653 Restraining {\it potentials} introduce repulsive potentials at the
172     surface of a sphere or other geometry. The solute and any explicit
173 gezelter 3665 solvent are therefore restrained inside the range defined by the
174     external potential. Often the potentials include a weak short-range
175     attraction to maintain the correct density at the boundary. Beglov
176     and Roux have also introduced a restraining boundary potential which
177     relaxes dynamically depending on the solute geometry and the force the
178     explicit system exerts on the shell.\cite{Beglov:1995fk}
179 gezelter 3653
180 gezelter 3665 Recently, Krilov {\it et al.} introduced a {\it flexible} boundary
181     model that uses a Lennard-Jones potential between the solvent
182     molecules and a boundary which is determined dynamically from the
183     position of the nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This
184     approach allows the confining potential to prevent solvent molecules
185     from migrating too far from the solute surface, while providing a weak
186     attractive force pulling the solvent molecules towards a fictitious
187     bulk solvent. Although this approach is appealing and has physical
188     motivation, nanoparticles do not deform far from their original
189     geometries even at temperatures which vaporize the nearby solvent. For
190     the systems like this, the flexible boundary model will be nearly
191 gezelter 3653 identical to a fixed-volume restraining potential.
192    
193 gezelter 3665 \subsection*{Hull methods}
194 gezelter 3653 The approach of Kohanoff, Caro, and Finnis is the most promising of
195     the methods for introducing both constant pressure and temperature
196     into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
197     This method is based on standard Langevin dynamics, but the Brownian
198     or random forces are allowed to act only on peripheral atoms and exert
199     force in a direction that is inward-facing relative to the facets of a
200     closed bounding surface. The statistical distribution of the random
201     forces are uniquely tied to the pressure in the external reservoir, so
202     the method can be shown to sample the isobaric-isothermal ensemble.
203     Kohanoff {\it et al.} used a Delaunay tessellation to generate a
204     bounding surface surrounding the outermost atoms in the simulated
205     system. This is not the only possible triangulated outer surface, but
206     guarantees that all of the random forces point inward towards the
207     cluster.
208    
209     In the following sections, we extend and generalize the approach of
210     Kohanoff, Caro, and Finnis. The new method, which we are calling the
211     ``Langevin Hull'' applies the external pressure, Langevin drag, and
212 gezelter 3667 random forces on the {\it facets of the hull} instead of the atomic
213     sites comprising the vertices of the hull. This allows us to decouple
214     the external pressure contribution from the drag and random force.
215     The methodology is introduced in section \ref{sec:meth}, tests on
216     crystalline nanoparticles, liquid clusters, and heterogeneous mixtures
217     are detailed in section \ref{sec:tests}. Section \ref{sec:discussion}
218     summarizes our findings.
219 gezelter 3653
220 gezelter 3640 \section{Methodology}
221 gezelter 3653 \label{sec:meth}
222 gezelter 3640
223 gezelter 3665 The Langevin Hull uses an external bath at a fixed constant pressure
224     ($P$) and temperature ($T$). This bath interacts only with the
225     objects on the exterior hull of the system. Defining the hull of the
226     simulation is done in a manner similar to the approach of Kohanoff,
227     Caro and Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous
228     configuration of the atoms in the system is considered as a point
229     cloud in three dimensional space. Delaunay triangulation is used to
230     find all facets between coplanar
231     neighbors.\cite{delaunay,springerlink:10.1007/BF00977785} In highly
232     symmetric point clouds, facets can contain many atoms, but in all but
233     the most symmetric of cases the facets are simple triangles in 3-space
234     that contain exactly three atoms.
235 gezelter 3640
236 gezelter 3652 The convex hull is the set of facets that have {\it no concave
237 gezelter 3665 corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
238     eliminates all facets on the interior of the point cloud, leaving only
239     those exposed to the bath. Sites on the convex hull are dynamic; as
240     molecules re-enter the cluster, all interactions between atoms on that
241     molecule and the external bath are removed. Since the edge is
242     determined dynamically as the simulation progresses, no {\it a priori}
243     geometry is defined. The pressure and temperature bath interacts only
244 gezelter 3660 with the atoms on the edge and not with atoms interior to the
245     simulation.
246 gezelter 3640
247 gezelter 3662 \begin{figure}
248 gezelter 3688 \includegraphics[width=\linewidth]{solvatedNano}
249 gezelter 3662 \caption{The external temperature and pressure bath interacts only
250     with those atoms on the convex hull (grey surface). The hull is
251 gezelter 3667 computed dynamically at each time step, and molecules can move
252     between the interior (Newtonian) region and the Langevin hull.}
253 gezelter 3662 \label{fig:hullSample}
254     \end{figure}
255    
256 gezelter 3665 Atomic sites in the interior of the simulation move under standard
257 gezelter 3660 Newtonian dynamics,
258 gezelter 3640 \begin{equation}
259 gezelter 3652 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
260     \label{eq:Newton}
261 gezelter 3640 \end{equation}
262 gezelter 3652 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
263     instantaneous velocity of site $i$ at time $t$, and $U$ is the total
264     potential energy. For atoms on the exterior of the cluster
265     (i.e. those that occupy one of the vertices of the convex hull), the
266     equation of motion is modified with an external force, ${\mathbf
267     F}_i^{\mathrm ext}$,
268 gezelter 3640 \begin{equation}
269 gezelter 3652 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
270 gezelter 3640 \end{equation}
271    
272 gezelter 3665 The external bath interacts indirectly with the atomic sites through
273     the intermediary of the hull facets. Since each vertex (or atom)
274     provides one corner of a triangular facet, the force on the facets are
275     divided equally to each vertex. However, each vertex can participate
276     in multiple facets, so the resultant force is a sum over all facets
277     $f$ containing vertex $i$:
278 gezelter 3640 \begin{equation}
279     {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
280     } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
281     F}_f^{\mathrm ext}
282     \end{equation}
283    
284 gezelter 3652 The external pressure bath applies a force to the facets of the convex
285     hull in direct proportion to the area of the facet, while the thermal
286 gezelter 3660 coupling depends on the solvent temperature, viscosity and the size
287     and shape of each facet. The thermal interactions are expressed as a
288     standard Langevin description of the forces,
289 gezelter 3640 \begin{equation}
290     \begin{array}{rclclcl}
291     {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
292     & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
293     \end{array}
294     \end{equation}
295 gezelter 3665 Here, $A_f$ and $\hat{n}_f$ are the area and (outward-facing) normal
296     vectors for facet $f$, respectively. ${\mathbf v}_f(t)$ is the
297     velocity of the facet centroid,
298 gezelter 3652 \begin{equation}
299     {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
300     \end{equation}
301 gezelter 3660 and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
302     depends on the geometry and surface area of facet $f$ and the
303     viscosity of the fluid. The resistance tensor is related to the
304     fluctuations of the random force, $\mathbf{R}(t)$, by the
305     fluctuation-dissipation theorem,
306 gezelter 3640 \begin{eqnarray}
307     \left< {\mathbf R}_f(t) \right> & = & 0 \\
308     \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
309 gezelter 3652 \Xi_f(t)\delta(t-t^\prime).
310     \label{eq:randomForce}
311 gezelter 3640 \end{eqnarray}
312    
313 gezelter 3665 Once the resistance tensor is known for a given facet, a stochastic
314 gezelter 3660 vector that has the properties in Eq. (\ref{eq:randomForce}) can be
315 gezelter 3665 calculated efficiently by carrying out a Cholesky decomposition to
316     obtain the square root matrix of the resistance tensor,
317 gezelter 3652 \begin{equation}
318     \Xi_f = {\bf S} {\bf S}^{T},
319     \label{eq:Cholesky}
320     \end{equation}
321     where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
322     vector with the statistics required for the random force can then be
323     obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
324     has elements chosen from a Gaussian distribution, such that:
325     \begin{equation}
326     \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
327     {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
328     \end{equation}
329     where $\delta t$ is the timestep in use during the simulation. The
330     random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
331     have the correct properties required by Eq. (\ref{eq:randomForce}).
332 gezelter 3640
333 gezelter 3660 Our treatment of the resistance tensor is approximate. $\Xi$ for a
334     rigid triangular plate would normally be treated as a $6 \times 6$
335 gezelter 3653 tensor that includes translational and rotational drag as well as
336 gezelter 3660 translational-rotational coupling. The computation of resistance
337 gezelter 3653 tensors for rigid bodies has been detailed
338 gezelter 3663 elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
339 gezelter 3653 but the standard approach involving bead approximations would be
340     prohibitively expensive if it were recomputed at each step in a
341     molecular dynamics simulation.
342    
343 gezelter 3665 Instead, we are utilizing an approximate resistance tensor obtained by
344     first constructing the Oseen tensor for the interaction of the
345     centroid of the facet ($f$) with each of the subfacets $\ell=1,2,3$,
346 gezelter 3653 \begin{equation}
347 gezelter 3665 T_{\ell f}=\frac{A_\ell}{8\pi\eta R_{\ell f}}\left(I +
348     \frac{\mathbf{R}_{\ell f}\mathbf{R}_{\ell f}^T}{R_{\ell f}^2}\right)
349 gezelter 3653 \end{equation}
350 gezelter 3665 Here, $A_\ell$ is the area of subfacet $\ell$ which is a triangle
351     containing two of the vertices of the facet along with the centroid.
352     $\mathbf{R}_{\ell f}$ is the vector between the centroid of facet $f$
353     and the centroid of sub-facet $\ell$, and $I$ is the ($3 \times 3$)
354     identity matrix. $\eta$ is the viscosity of the external bath.
355 gezelter 3653
356     \begin{figure}
357     \includegraphics[width=\linewidth]{hydro}
358 gezelter 3660 \caption{The resistance tensor $\Xi$ for a facet comprising sites $i$,
359     $j$, and $k$ is constructed using Oseen tensor contributions between
360     the centoid of the facet $f$ and each of the sub-facets ($i,f,j$),
361     ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets are
362     located at $1$, $2$, and $3$, and the area of each sub-facet is
363 gezelter 3653 easily computed using half the cross product of two of the edges.}
364     \label{hydro}
365     \end{figure}
366    
367 gezelter 3665 The tensors for each of the sub-facets are added together, and the
368     resulting matrix is inverted to give a $3 \times 3$ resistance tensor
369     for translations of the triangular facet,
370 gezelter 3653 \begin{equation}
371     \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
372     \end{equation}
373 gezelter 3667 Note that this treatment ignores rotations (and
374 gezelter 3660 translational-rotational coupling) of the facet. In compact systems,
375     the facets stay relatively fixed in orientation between
376     configurations, so this appears to be a reasonably good approximation.
377    
378 gezelter 3652 We have implemented this method by extending the Langevin dynamics
379 gezelter 3665 integrator in our code, OpenMD.\cite{Meineke2005,openmd} At each
380     molecular dynamics time step, the following process is carried out:
381     \begin{enumerate}
382     \item The standard inter-atomic forces ($\nabla_iU$) are computed.
383 gezelter 3667 \item Delaunay triangulation is carried out using the current atomic
384 gezelter 3665 configuration.
385     \item The convex hull is computed and facets are identified.
386     \item For each facet:
387     \begin{itemize}
388     \item[a.] The force from the pressure bath ($-PA_f\hat{n}_f$) is
389     computed.
390     \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
391     viscosity ($\eta$) of the bath.
392     \item[c.] Facet drag ($-\Xi_f(t) \mathbf{v}_f(t)$) forces are
393     computed.
394     \item[d.] Random forces ($\mathbf{R}_f(t)$) are computed using the
395     resistance tensor and the temperature ($T$) of the bath.
396     \end{itemize}
397     \item The facet forces are divided equally among the vertex atoms.
398     \item Atomic positions and velocities are propagated.
399     \end{enumerate}
400     The Delaunay triangulation and computation of the convex hull are done
401     using calls to the qhull library.\cite{Qhull} There is a minimal
402     penalty for computing the convex hull and resistance tensors at each
403     step in the molecular dynamics simulation (roughly 0.02 $\times$ cost
404     of a single force evaluation), and the convex hull is remarkably easy
405     to parallelize on distributed memory machines (see Appendix A).
406 gezelter 3652
407 gezelter 3640 \section{Tests \& Applications}
408 gezelter 3653 \label{sec:tests}
409 gezelter 3640
410 gezelter 3663 To test the new method, we have carried out simulations using the
411     Langevin Hull on: 1) a crystalline system (gold nanoparticles), 2) a
412 gezelter 3665 liquid droplet (SPC/E water),\cite{Berendsen1987} and 3) a
413     heterogeneous mixture (gold nanoparticles in a water droplet). In each
414     case, we have computed properties that depend on the external applied
415     pressure. Of particular interest for the single-phase systems is the
416     isothermal compressibility,
417 gezelter 3660 \begin{equation}
418     \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
419     )_{T}.
420     \label{eq:BM}
421     \end{equation}
422    
423     One problem with eliminating periodic boundary conditions and
424     simulation boxes is that the volume of a three-dimensional point cloud
425     is not well-defined. In order to compute the compressibility of a
426     bulk material, we make an assumption that the number density, $\rho =
427 gezelter 3665 \frac{N}{V}$, is uniform within some region of the point cloud. The
428 gezelter 3660 compressibility can then be expressed in terms of the average number
429     of particles in that region,
430     \begin{equation}
431 gezelter 3665 \kappa_{T} = -\frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
432 gezelter 3660 )_{T}
433     \label{eq:BMN}
434     \end{equation}
435 gezelter 3663 The region we used is a spherical volume of 10 \AA\ radius centered in
436     the middle of the cluster. $N$ is the average number of molecules
437     found within this region throughout a given simulation. The geometry
438     and size of the region is arbitrary, and any bulk-like portion of the
439 gezelter 3665 cluster can be used to compute the compressibility.
440 gezelter 3660
441 gezelter 3665 One might assume that the volume of the convex hull could simply be
442     taken as the system volume $V$ in the compressibility expression
443     (Eq. \ref{eq:BM}), but this has implications at lower pressures (which
444     are explored in detail in the section on water droplets).
445 gezelter 3660
446 gezelter 3663 The metallic force field in use for the gold nanoparticles is the
447     quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} In all
448     simulations involving point charges, we utilized damped shifted-force
449     (DSF) electrostatics\cite{Fennell06} which is a variant of the Wolf
450     summation\cite{wolf:8254} that has been shown to provide good forces
451     and torques on molecular models for water in a computationally
452     efficient manner.\cite{Fennell06} The damping parameter ($\alpha$) was
453     set to 0.18 \AA$^{-1}$, and the cutoff radius was set to 12 \AA. The
454     Spohr potential was adopted in depicting the interaction between metal
455     atoms and the SPC/E water molecules.\cite{ISI:000167766600035}
456    
457 gezelter 3667 \subsection{Compressibility of gold nanoparticles}
458 gezelter 3640
459 gezelter 3678 The compressibility (and its inverse, the bulk modulus) is well-known
460     for gold, and is captured well by the embedded atom method
461     (EAM)~\cite{PhysRevB.33.7983} potential
462     and related multi-body force fields. In particular, the quantum
463     Sutton-Chen potential gets nearly quantitative agreement with the
464     experimental bulk modulus values, and makes a good first test of how
465     the Langevin Hull will perform at large applied pressures.
466 gezelter 3663
467 gezelter 3678 The Sutton-Chen (SC) potentials are based on a model of a metal which
468     treats the nuclei and core electrons as pseudo-atoms embedded in the
469     electron density due to the valence electrons on all of the other
470     atoms in the system.\cite{Chen90} The SC potential has a simple form that closely
471     resembles the Lennard Jones potential,
472     \begin{equation}
473     \label{eq:SCP1}
474     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] ,
475     \end{equation}
476     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
477     \begin{equation}
478     \label{eq:SCP2}
479     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}}.
480     \end{equation}
481     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
482     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
483     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
484     the interactions between the valence electrons and the cores of the
485     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
486     scale, $c_i$ scales the attractive portion of the potential relative
487     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
488     that assures a dimensionless form for $\rho$. These parameters are
489     tuned to various experimental properties such as the density, cohesive
490     energy, and elastic moduli for FCC transition metals. The quantum
491     Sutton-Chen (QSC) formulation matches these properties while including
492     zero-point quantum corrections for different transition
493     metals.\cite{PhysRevB.59.3527}
494    
495     In bulk gold, the experimentally-measured value for the bulk modulus
496     is 180.32 GPa, while previous calculations on the QSC potential in
497     periodic-boundary simulations of the bulk have yielded values of
498     175.53 GPa.\cite{XXX} Using the same force field, we have performed a
499     series of relatively short (200 ps) simulations on 40 \r{A} radius
500     nanoparticles under the Langevin Hull at a variety of applied
501     pressures ranging from 0 GPa to XXX. We obtain a value of 177.547 GPa
502     for the bulk modulus for gold using this echnique.
503    
504 gezelter 3640 \begin{figure}
505 gezelter 3678 \includegraphics[width=\linewidth]{stacked}
506     \caption{The response of the internal pressure and temperature of gold
507     nanoparticles when first placed in the Langevin Hull
508     ($T_\mathrm{bath}$ = 300K, $P_\mathrm{bath}$ = 4 GPa), starting
509 gezelter 3665 from initial conditions that were far from the bath pressure and
510 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).}
511 gezelter 3640 \label{pressureResponse}
512     \end{figure}
513    
514     \begin{equation}
515     \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
516     P}\right)
517     \end{equation}
518    
519     \subsection{Compressibility of SPC/E water clusters}
520    
521 gezelter 3660 Prior molecular dynamics simulations on SPC/E water (both in
522     NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
523     ensembles) have yielded values for the isothermal compressibility that
524     agree well with experiment.\cite{Fine1973} The results of two
525     different approaches for computing the isothermal compressibility from
526     Langevin Hull simulations for pressures between 1 and 6500 atm are
527     shown in Fig. \ref{fig:compWater} along with compressibility values
528     obtained from both other SPC/E simulations and experiment.
529     Compressibility values from all references are for applied pressures
530     within the range 1 - 1000 atm.
531 kstocke1 3649
532 gezelter 3640 \begin{figure}
533 gezelter 3659 \includegraphics[width=\linewidth]{new_isothermalN}
534 kstocke1 3649 \caption{Compressibility of SPC/E water}
535 gezelter 3660 \label{fig:compWater}
536 gezelter 3640 \end{figure}
537    
538 gezelter 3660 Isothermal compressibility values calculated using the number density
539     (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
540     and previous simulation work throughout the 1 - 1000 atm pressure
541     regime. Compressibilities computed using the Hull volume, however,
542     deviate dramatically from the experimental values at low applied
543     pressures. The reason for this deviation is quite simple; at low
544     applied pressures, the liquid is in equilibrium with a vapor phase,
545     and it is entirely possible for one (or a few) molecules to drift away
546     from the liquid cluster (see Fig. \ref{fig:coneOfShame}). At low
547     pressures, the restoring forces on the facets are very gentle, and
548     this means that the hulls often take on relatively distorted
549     geometries which include large volumes of empty space.
550 kstocke1 3649
551 gezelter 3660 \begin{figure}
552 gezelter 3688 \includegraphics[width=\linewidth]{coneOfShame}
553 gezelter 3660 \caption{At low pressures, the liquid is in equilibrium with the vapor
554     phase, and isolated molecules can detach from the liquid droplet.
555 gezelter 3665 This is expected behavior, but the volume of the convex hull
556     includes large regions of empty space. For this reason,
557 gezelter 3662 compressibilities are computed using local number densities rather
558     than hull volumes.}
559 gezelter 3660 \label{fig:coneOfShame}
560     \end{figure}
561 kstocke1 3649
562 gezelter 3665 At higher pressures, the equilibrium strongly favors the liquid phase,
563     and the hull geometries are much more compact. Because of the
564     liquid-vapor effect on the convex hull, the regional number density
565     approach (Eq. \ref{eq:BMN}) provides more reliable estimates of the
566 gezelter 3667 compressibility.
567 kstocke1 3649
568 gezelter 3665 In both the traditional compressibility formula (Eq. \ref{eq:BM}) and
569     the number density version (Eq. \ref{eq:BMN}), multiple simulations at
570     different pressures must be done to compute the first derivatives. It
571     is also possible to compute the compressibility using the fluctuation
572     dissipation theorem using either fluctuations in the
573     volume,\cite{Debenedetti1986},
574 kstocke1 3649 \begin{equation}
575 gezelter 3665 \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle
576     V \right \rangle ^{2}}{V \, k_{B} \, T},
577 kstocke1 3649 \end{equation}
578 gezelter 3665 or, equivalently, fluctuations in the number of molecules within the
579     fixed region,
580     \begin{equation}
581     \kappa_{T} = \frac{\left \langle N^{2} \right \rangle - \left \langle
582     N \right \rangle ^{2}}{N \, k_{B} \, T},
583     \end{equation}
584     Thus, the compressibility of each simulation can be calculated
585     entirely independently from all other trajectories. However, the
586     resulting compressibilities were still as much as an order of
587 gezelter 3667 magnitude larger than the reference values. However, compressibility
588 gezelter 3665 calculation that relies on the hull volume will suffer these effects.
589     WE NEED MORE HERE.
590 kstocke1 3649
591     \subsection{Molecular orientation distribution at cluster boundary}
592    
593 gezelter 3665 In order for non-periodic boundary conditions to be widely applicable,
594     they must be constructed in such a way that they allow a finite system
595 gezelter 3667 to replicate the properties of the bulk. Early non-periodic
596     simulation methods (e.g. hydrophobic boundary potentials) induced
597     spurious orientational correlations deep within the simulated
598     system.\cite{Lee1984,Belch1985} This behavior spawned many methods for
599     fixing and characterizing the effects of artifical boundaries
600     including methods which fix the orientations of a set of edge
601     molecules.\cite{Warshel1978,King1989}
602 kstocke1 3649
603 gezelter 3667 As described above, the Langevin Hull does not require that the
604     orientation of molecules be fixed, nor does it utilize an explicitly
605     hydrophobic boundary, orientational constraint or radial constraint.
606     Therefore, the orientational correlations of the molecules in a water
607     cluster are of particular interest in testing this method. Ideally,
608     the water molecules on the surface of the cluster will have enough
609     mobility into and out of the center of the cluster to maintain a
610     bulk-like orientational distribution in the absence of orientational
611     and radial constraints. However, since the number of hydrogen bonding
612     partners available to molecules on the exterior are limited, it is
613     likely that there will be some effective hydrophobicity of the hull.
614 kstocke1 3649
615 gezelter 3667 To determine the extent of these effects demonstrated by the Langevin
616     Hull, we examined the orientationations exhibited by SPC/E water in a
617     cluster of 1372 molecules at 300 K and at pressures ranging from 1 -
618     1000 atm. The orientational angle of a water molecule is described
619 kstocke1 3649 \begin{equation}
620 gezelter 3640 \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
621     \end{equation}
622 gezelter 3667 where $\vec{r}_{i}$ is the vector between molecule {\it i}'s center of
623     mass and the cluster center of mass and $\vec{\mu}_{i}$ is the vector
624 gezelter 3678 bisecting the H-O-H angle of molecule {\it i} Bulk-like distributions
625     will result in $\langle \cos \theta \rangle$ values close to zero. If
626     the hull exhibits an overabundance of externally-oriented oxygen sites
627     the average orientation will be negative, while dangling hydrogen
628     sites will result in positive average orientations.
629 kstocke1 3649
630 gezelter 3667 Fig. \ref{fig:pAngle} shows the distribution of $\cos{\theta}$ values
631     for molecules in the interior of the cluster (squares) and for
632     molecules included in the convex hull (circles).
633 kstocke1 3649 \begin{figure}
634 gezelter 3640 \includegraphics[width=\linewidth]{pAngle}
635 gezelter 3667 \caption{Distribution of $\cos{\theta}$ values for molecules on the
636     interior of the cluster (squares) and for those participating in the
637     convex hull (circles) at a variety of pressures. The Langevin hull
638     exhibits minor dewetting behavior with exposed oxygen sites on the
639     hull water molecules. The orientational preference for exposed
640     oxygen appears to be independent of applied pressure. }
641     \label{fig:pAngle}
642 gezelter 3640 \end{figure}
643    
644 gezelter 3667 As expected, interior molecules (those not included in the convex
645     hull) maintain a bulk-like structure with a uniform distribution of
646     orientations. Molecules included in the convex hull show a slight
647     preference for values of $\cos{\theta} < 0.$ These values correspond
648     to molecules with oxygen directed toward the exterior of the cluster,
649     forming a dangling hydrogen bond acceptor site.
650 gezelter 3640
651 gezelter 3667 In the absence of an electrostatic contribution from the exterior
652     bath, the orientational distribution of water molecules included in
653     the Langevin Hull will slightly resemble the distribution at a neat
654     water liquid/vapor interface. Previous molecular dynamics simulations
655     of SPC/E water \cite{Taylor1996} have shown that molecules at the
656     liquid/vapor interface favor an orientation where one hydrogen
657     protrudes from the liquid phase. This behavior is demonstrated by
658     experiments \cite{Du1994} \cite{Scatena2001} showing that
659     approximately one-quarter of water molecules at the liquid/vapor
660     interface form dangling hydrogen bonds. The negligible preference
661     shown in these cluster simulations could be removed through the
662     introduction of an implicit solvent model, which would provide the
663     missing electrostatic interactions between the cluster molecules and
664     the surrounding temperature/pressure bath.
665 kstocke1 3649
666 gezelter 3667 The orientational preference exhibited by hull molecules in the
667     Langevin hull is significantly weaker than the preference caused by an
668     explicit hydrophobic bounding potential. Additionally, the Langevin
669     Hull does not require that the orientation of any molecules be fixed
670     in order to maintain bulk-like structure, even at the cluster surface.
671 kstocke1 3649
672 gezelter 3640 \subsection{Heterogeneous nanoparticle / water mixtures}
673    
674 gezelter 3665 \section{Discussion}
675     \label{sec:discussion}
676 gezelter 3640
677 gezelter 3667 The Langevin Hull samples the isobaric-isothermal ensemble for
678     non-periodic systems by coupling the system to an bath characterized
679     by pressure, temperature, and solvent viscosity. This enables the
680     study of heterogeneous systems composed of materials of significantly
681     different compressibilities. Because the boundary is dynamically
682     determined during the simulation and the molecules interacting with
683     the boundary can change, the method and has minimal perturbations on
684     the behavior of molecules at the edges of the simulation. Further
685     work on this method will involve implicit electrostatics at the
686     boundary (which is missing in the current implementation) as well as
687     more sophisticated treatments of the surface geometry (alpha
688     shapes\cite{EDELSBRUNNER:1994oq,EDELSBRUNNER:1995cj} and Tight
689     Cocone\cite{Dey:2003ts}). The non-convex hull geometries are
690     significantly more expensive ($\mathcal{O}(N^2)$) than the convex hull
691     ($\mathcal{O}(N \log N)$), but would enable the use of hull volumes
692     directly in computing the compressibility of the sample.
693    
694 gezelter 3663 \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
695 gezelter 3640
696 gezelter 3666 In order to use the Langevin Hull for simulations on parallel
697     computers, one of the more difficult tasks is to compute the bounding
698     surface, facets, and resistance tensors when the processors have
699     incomplete information about the entire system's topology. Most
700     parallel decomposition methods assign primary responsibility for the
701     motion of an atomic site to a single processor, and we can exploit
702     this to efficiently compute the convex hull for the entire system.
703    
704 gezelter 3667 The basic idea involves splitting the point cloud into
705     spatially-overlapping subsets and computing the convex hulls for each
706     of the subsets. The points on the convex hull of the entire system
707     are all present on at least one of the subset hulls. The algorithm
708     works as follows:
709 gezelter 3666 \begin{enumerate}
710     \item Each processor computes the convex hull for its own atomic sites
711 gezelter 3668 (left panel in Fig. \ref{fig:parallel}).
712 gezelter 3684 \item The Hull vertices from each processor are communicated to all of
713 gezelter 3666 the processors, and each processor assembles a complete list of hull
714     sites (this is much smaller than the original number of points in
715     the point cloud).
716 gezelter 3668 \item Each processor computes the global convex hull (right panel in
717 gezelter 3667 Fig. \ref{fig:parallel}) using only those points that are the union
718     of sites gathered from all of the subset hulls. Delaunay
719     triangulation is then done to obtain the facets of the global hull.
720 gezelter 3666 \end{enumerate}
721    
722     \begin{figure}
723 gezelter 3668 \includegraphics[width=\linewidth]{parallel}
724 gezelter 3666 \caption{When the sites are distributed among many nodes for parallel
725     computation, the processors first compute the convex hulls for their
726 gezelter 3668 own sites (dashed lines in left panel). The positions of the sites
727 gezelter 3669 that make up the subset hulls are then communicated to all
728 gezelter 3684 processors (middle panel). The convex hull of the system (solid line in
729     right panel) is the convex hull of the points on the union of the subset
730     hulls.}
731 gezelter 3668 \label{fig:parallel}
732 gezelter 3666 \end{figure}
733    
734     The individual hull operations scale with
735 gezelter 3667 $\mathcal{O}(\frac{n}{p}\log\frac{n}{p})$ where $n$ is the total
736     number of sites, and $p$ is the number of processors. These local
737     hull operations create a set of $p$ hulls each with approximately
738     $\frac{n}{3pr}$ sites (for a cluster of radius $r$). The worst-case
739     communication cost for using a ``gather'' operation to distribute this
740     information to all processors is $\mathcal{O}( \alpha (p-1) + \frac{n
741     \beta (p-1)}{3 r p^2})$, while the final computation of the system
742     hull scales as $\mathcal{O}(\frac{n}{3r}\log\frac{n}{3r})$.
743 gezelter 3666
744 gezelter 3667 For a large number of atoms on a moderately parallel machine, the
745     total costs are dominated by the computations of the individual hulls,
746     and communication of these hulls to so the Langevin hull sees roughly
747     linear speed-up with increasing processor counts.
748    
749 gezelter 3663 \section*{Acknowledgments}
750 gezelter 3640 Support for this project was provided by the
751     National Science Foundation under grant CHE-0848243. Computational
752     time was provided by the Center for Research Computing (CRC) at the
753     University of Notre Dame.
754    
755 gezelter 3685 Molecular graphics images were produced using the UCSF Chimera package from
756     the Resource for Biocomputing, Visualization, and Informatics at the
757     University of California, San Francisco (supported by NIH P41 RR001081).
758 gezelter 3640 \newpage
759    
760     \bibliography{langevinHull}
761    
762     \end{doublespace}
763     \end{document}