ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevinHull/langevinHull.tex
(Generate patch)

Comparing trunk/langevinHull/langevinHull.tex (file contents):
Revision 3652 by gezelter, Mon Oct 18 18:27:24 2010 UTC vs.
Revision 3678 by gezelter, Mon Nov 8 20:11:36 2010 UTC

# Line 18 | Line 18
18   \setlength{\belowcaptionskip}{30 pt}
19  
20   \bibpunct{[}{]}{,}{s}{}{;}
21 < \bibliographystyle{aip}
21 > \bibliographystyle{achemso}
22  
23   \begin{document}
24  
# Line 39 | Line 39 | Notre Dame, Indiana 46556}
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
42 >  hull surrounding the system.  A Langevin thermostat is also applied
43 >  to facets of the hull to mimic contact with an external heat
44 >  bath. This new method, the ``Langevin Hull'', performs better than
45 >  traditional affine transform methods for systems containing
46 >  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.
51 >  metal nanoparticles, nanoparticles in an explicit solvent, as well
52 >  as clusters of liquid water. The predicted mechanical properties of
53 >  these systems are in good agreement with experimental data and
54 >  previous simulation work.
55   \end{abstract}
56  
57   \newpage
# Line 66 | Line 66 | of an isobaric-isothermal (NPT) ensemble attempt to ma
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
69 > 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 coordinate
84 < transformation to adjust the box volume.
83 > Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize scaled
84 > coordinate transformation to adjust the box volume.  As long as the
85 > 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  
90 < As long as the material in the simulation box is essentially a bulk
91 < liquid which has a relatively uniform compressibility, the standard
92 < approach provides an excellent way of adjusting the volume of the
93 < system and applying pressure directly via the interactions between
94 < atomic sites.  
90 <
91 < The problem with these approaches becomes apparent when the material
92 < being simulated is an inhomogeneous mixture in which portions of the
93 < simulation box are incompressible relative to other portions.
94 < Examples include simulations of metallic nanoparticles in liquid
95 < environments, proteins at interfaces, as well as other multi-phase or
90 > 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   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 collisions in the incompressible region.
97 > 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  
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.}
103 > \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   \label{affineScale}
111   \end{figure}
112  
113 < Additionally, one may often wish to simulate explicitly non-periodic
114 < systems, and the constraint that a periodic box must be used to
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 requires either 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 < Explicitly non-periodic systems
123 > \subsection*{Boundary Methods}
124 > There have been a number of approaches to handle simulations of
125 > explicitly non-periodic systems that focus on constant or
126 > nearly-constant {\it volume} conditions while maintaining bulk-like
127 > behavior.  Berkowitz and McCammon introduced a stochastic (Langevin)
128 > boundary layer inside a region of fixed molecules which effectively
129 > enforces constant temperature and volume (NVT)
130 > conditions.\cite{Berkowitz1982} In this approach, the stochastic and
131 > fixed regions were defined relative to a central atom.  Brooks and
132 > Karplus extended this method to include deformable stochastic
133 > boundaries.\cite{iii:6312} The stochastic boundary approach has been
134 > used widely for protein simulations. [CITATIONS NEEDED]
135  
136 < Elastic Bag
136 > The electrostatic and dispersive behavior near the boundary has long
137 > been a cause for concern when performing simulations of explicitly
138 > non-periodic systems.  Early work led to the surface constrained soft
139 > sphere dipole model (SCSSD)\cite{Warshel1978} in which the surface
140 > molecules are fixed in a random orientation representative of the bulk
141 > solvent structural properties. Belch {\it et al.}\cite{Belch1985}
142 > simulated clusters of TIPS2 water surrounded by a hydrophobic bounding
143 > potential. The spherical hydrophobic boundary induced dangling
144 > hydrogen bonds at the surface that propagated deep into the cluster,
145 > affecting most of molecules in the simulation.  This result echoes an
146 > earlier study which showed that an extended planar hydrophobic surface
147 > caused orientational preference at the surface which extended
148 > relatively deep (7 \r{A}) into the liquid simulation
149 > cell.\cite{Lee1984} The surface constrained all-atom solvent (SCAAS)
150 > model \cite{King1989} improved upon its SCSSD predecessor. The SCAAS
151 > model utilizes a polarization constraint which is applied to the
152 > surface molecules to maintain bulk-like structure at the cluster
153 > surface. A radial constraint is used to maintain the desired bulk
154 > density of the liquid. Both constraint forces are applied only to a
155 > pre-determined number of the outermost molecules.
156  
157 < Spherical Boundary approaches
157 > Beglov and Roux have developed a boundary model in which the hard
158 > sphere boundary has a radius that varies with the instantaneous
159 > configuration of the solute (and solvent) molecules.\cite{beglov:9050}
160 > This model contains a clear pressure and surface tension contribution
161 > to the free energy which XXX.
162  
163 < \section{Methodology}
163 > \subsection*{Restraining Potentials}
164 > Restraining {\it potentials} introduce repulsive potentials at the
165 > surface of a sphere or other geometry.  The solute and any explicit
166 > solvent are therefore restrained inside the range defined by the
167 > external potential.  Often the potentials include a weak short-range
168 > attraction to maintain the correct density at the boundary.  Beglov
169 > and Roux have also introduced a restraining boundary potential which
170 > relaxes dynamically depending on the solute geometry and the force the
171 > explicit system exerts on the shell.\cite{Beglov:1995fk}
172  
173 < We have developed a new method which uses a constant pressure and
174 < temperature bath.  This bath interacts only with the objects that are
175 < currently at the edge of the system.  Since the edge is determined
176 < dynamically as the simulation progresses, no {\it a priori} geometry
177 < is defined.  The pressure and temperature bath interacts {\it
178 <  directly} with the atoms on the edge and not with atoms interior to
179 < the simulation.  This means that there are no affine transforms
180 < required.  There are also no fictitious particles or bounding
181 < potentials used in this approach.
173 > Recently, Krilov {\it et al.} introduced a {\it flexible} boundary
174 > model that uses a Lennard-Jones potential between the solvent
175 > molecules and a boundary which is determined dynamically from the
176 > position of the nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This
177 > approach allows the confining potential to prevent solvent molecules
178 > from migrating too far from the solute surface, while providing a weak
179 > attractive force pulling the solvent molecules towards a fictitious
180 > bulk solvent.  Although this approach is appealing and has physical
181 > motivation, nanoparticles do not deform far from their original
182 > geometries even at temperatures which vaporize the nearby solvent. For
183 > the systems like this, the flexible boundary model will be nearly
184 > identical to a fixed-volume restraining potential.
185  
186 < The basics of the method are as follows. The simulation starts as a
187 < collection of atomic locations in three dimensions (a point cloud).
188 < Delaunay triangulation is used to find all facets between coplanar
189 < neighbors.  In highly symmetric point clouds, facets can contain many
190 < atoms, but in all but the most symmetric of cases one might experience
191 < in a molecular dynamics simulation, the facets are simple triangles in
192 < 3-space that contain exactly three atoms.  
186 > \subsection*{Hull methods}
187 > The approach of Kohanoff, Caro, and Finnis is the most promising of
188 > the methods for introducing both constant pressure and temperature
189 > into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
190 > This method is based on standard Langevin dynamics, but the Brownian
191 > or random forces are allowed to act only on peripheral atoms and exert
192 > force in a direction that is inward-facing relative to the facets of a
193 > closed bounding surface.  The statistical distribution of the random
194 > forces are uniquely tied to the pressure in the external reservoir, so
195 > the method can be shown to sample the isobaric-isothermal ensemble.
196 > Kohanoff {\it et al.} used a Delaunay tessellation to generate a
197 > bounding surface surrounding the outermost atoms in the simulated
198 > system.  This is not the only possible triangulated outer surface, but
199 > guarantees that all of the random forces point inward towards the
200 > cluster.
201 >
202 > In the following sections, we extend and generalize the approach of
203 > Kohanoff, Caro, and Finnis. The new method, which we are calling the
204 > ``Langevin Hull'' applies the external pressure, Langevin drag, and
205 > random forces on the {\it facets of the hull} instead of the atomic
206 > sites comprising the vertices of the hull.  This allows us to decouple
207 > the external pressure contribution from the drag and random force.
208 > The methodology is introduced in section \ref{sec:meth}, tests on
209 > crystalline nanoparticles, liquid clusters, and heterogeneous mixtures
210 > are detailed in section \ref{sec:tests}.  Section \ref{sec:discussion}
211 > summarizes our findings.
212  
213 + \section{Methodology}
214 + \label{sec:meth}
215 +
216 + The Langevin Hull uses an external bath at a fixed constant pressure
217 + ($P$) and temperature ($T$).  This bath interacts only with the
218 + objects on the exterior hull of the system.  Defining the hull of the
219 + simulation is done in a manner similar to the approach of Kohanoff,
220 + Caro and Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous
221 + configuration of the atoms in the system is considered as a point
222 + cloud in three dimensional space.  Delaunay triangulation is used to
223 + find all facets between coplanar
224 + neighbors.\cite{delaunay,springerlink:10.1007/BF00977785}  In highly
225 + symmetric point clouds, facets can contain many atoms, but in all but
226 + the most symmetric of cases the facets are simple triangles in 3-space
227 + that contain exactly three atoms.
228 +
229   The convex hull is the set of facets that have {\it no concave
230 <  corners} at an atomic site.  This eliminates all facets on the
231 < interior of the point cloud, leaving only those exposed to the
232 < bath. Sites on the convex hull are dynamic. As molecules re-enter the
233 < cluster, all interactions between atoms on that molecule and the
234 < external bath are removed.
230 >  corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
231 > eliminates all facets on the interior of the point cloud, leaving only
232 > those exposed to the bath. Sites on the convex hull are dynamic; as
233 > molecules re-enter the cluster, all interactions between atoms on that
234 > molecule and the external bath are removed.  Since the edge is
235 > determined dynamically as the simulation progresses, no {\it a priori}
236 > geometry is defined. The pressure and temperature bath interacts only
237 > with the atoms on the edge and not with atoms interior to the
238 > simulation.
239  
240 < For atomic sites in the interior of the point cloud, the equations of
241 < motion are simple Newtonian dynamics,
240 > \begin{figure}
241 > \includegraphics[width=\linewidth]{hullSample}
242 > \caption{The external temperature and pressure bath interacts only
243 >  with those atoms on the convex hull (grey surface).  The hull is
244 >  computed dynamically at each time step, and molecules can move
245 >  between the interior (Newtonian) region and the Langevin hull.}
246 > \label{fig:hullSample}
247 > \end{figure}
248 >
249 > Atomic sites in the interior of the simulation move under standard
250 > Newtonian dynamics,
251   \begin{equation}
252   m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
253   \label{eq:Newton}
# Line 163 | Line 262 | The external bath interacts directly with the facets o
262   m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
263   \end{equation}
264  
265 < The external bath interacts directly with the facets of the convex
266 < hull.  Since each vertex (or atom) provides one corner of a triangular
267 < facet, the force on the facets are divided equally to each vertex.
268 < However, each vertex can participate in multiple facets, so the resultant
269 < force is a sum over all facets $f$ containing vertex $i$:
265 > The external bath interacts indirectly with the atomic sites through
266 > the intermediary of the hull facets.  Since each vertex (or atom)
267 > provides one corner of a triangular facet, the force on the facets are
268 > divided equally to each vertex.  However, each vertex can participate
269 > in multiple facets, so the resultant force is a sum over all facets
270 > $f$ containing vertex $i$:
271   \begin{equation}
272   {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
273      } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\  {\mathbf
# Line 176 | Line 276 | coupling depends on the solvent temperature, friction
276  
277   The external pressure bath applies a force to the facets of the convex
278   hull in direct proportion to the area of the facet, while the thermal
279 < coupling depends on the solvent temperature, friction and the size and
280 < shape of each facet. The thermal interactions are expressed as a
281 < typical Langevin description of the forces,
279 > coupling depends on the solvent temperature, viscosity and the size
280 > and shape of each facet. The thermal interactions are expressed as a
281 > standard Langevin description of the forces,
282   \begin{equation}
283   \begin{array}{rclclcl}
284   {\mathbf F}_f^{\text{ext}} & = &  \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
285   & = &  -\hat{n}_f P A_f  & - & \Xi_f(t) {\mathbf v}_f(t)  & + & {\mathbf R}_f(t)
286   \end{array}
287   \end{equation}
288 < Here, $P$ is the external pressure, $A_f$ and $\hat{n}_f$ are the area
289 < and normal vectors for facet $f$, respectively.  ${\mathbf v}_f(t)$ is
290 < the velocity of the facet,
288 > Here, $A_f$ and $\hat{n}_f$ are the area and (outward-facing) normal
289 > vectors for facet $f$, respectively.  ${\mathbf v}_f(t)$ is the
290 > velocity of the facet centroid,
291   \begin{equation}
292   {\mathbf v}_f(t) =  \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
293   \end{equation}
294 < and $\Xi_f(t)$ is a ($3 \times 3$) hydrodynamic tensor that depends on
295 < the geometry and surface area of facet $f$ and the viscosity of the
296 < fluid (See Appendix A).  The hydrodynamic tensor is related to the
294 > and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
295 > depends on the geometry and surface area of facet $f$ and the
296 > viscosity of the fluid.  The resistance tensor is related to the
297   fluctuations of the random force, $\mathbf{R}(t)$, by the
298   fluctuation-dissipation theorem,
299   \begin{eqnarray}
# Line 203 | Line 303 | Once the hydrodynamic tensor is known for a given face
303   \label{eq:randomForce}
304   \end{eqnarray}
305  
306 < Once the hydrodynamic tensor is known for a given facet (see Appendix
307 < A) obtaining a stochastic vector that has the properties in
308 < Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
309 < one-time Cholesky decomposition to obtain the square root matrix of
210 < the resistance tensor,
306 > Once the resistance tensor is known for a given facet, a stochastic
307 > vector that has the properties in Eq. (\ref{eq:randomForce}) can be
308 > calculated efficiently by carrying out a Cholesky decomposition to
309 > obtain the square root matrix of the resistance tensor,
310   \begin{equation}
311   \Xi_f = {\bf S} {\bf S}^{T},
312   \label{eq:Cholesky}
# Line 224 | Line 323 | We have implemented this method by extending the Lange
323   random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
324   have the correct properties required by Eq. (\ref{eq:randomForce}).
325  
326 + Our treatment of the resistance tensor is approximate.  $\Xi$ for a
327 + rigid triangular plate would normally be treated as a $6 \times 6$
328 + tensor that includes translational and rotational drag as well as
329 + translational-rotational coupling. The computation of resistance
330 + tensors for rigid bodies has been detailed
331 + elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
332 + but the standard approach involving bead approximations would be
333 + prohibitively expensive if it were recomputed at each step in a
334 + molecular dynamics simulation.
335 +
336 + Instead, we are utilizing an approximate resistance tensor obtained by
337 + first constructing the Oseen tensor for the interaction of the
338 + centroid of the facet ($f$) with each of the subfacets $\ell=1,2,3$,
339 + \begin{equation}
340 + T_{\ell f}=\frac{A_\ell}{8\pi\eta R_{\ell f}}\left(I +
341 +  \frac{\mathbf{R}_{\ell f}\mathbf{R}_{\ell f}^T}{R_{\ell f}^2}\right)
342 + \end{equation}
343 + Here, $A_\ell$ is the area of subfacet $\ell$ which is a triangle
344 + containing two of the vertices of the facet along with the centroid.
345 + $\mathbf{R}_{\ell f}$ is the vector between the centroid of facet $f$
346 + and the centroid of sub-facet $\ell$, and $I$ is the ($3 \times 3$)
347 + identity matrix.  $\eta$ is the viscosity of the external bath.
348 +
349 + \begin{figure}
350 + \includegraphics[width=\linewidth]{hydro}
351 + \caption{The resistance tensor $\Xi$ for a facet comprising sites $i$,
352 +  $j$, and $k$ is constructed using Oseen tensor contributions between
353 +  the centoid of the facet $f$ and each of the sub-facets ($i,f,j$),
354 +  ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets are
355 +  located at $1$, $2$, and $3$, and the area of each sub-facet is
356 +  easily computed using half the cross product of two of the edges.}
357 + \label{hydro}
358 + \end{figure}
359 +
360 + The tensors for each of the sub-facets are added together, and the
361 + resulting matrix is inverted to give a $3 \times 3$ resistance tensor
362 + for translations of the triangular facet,
363 + \begin{equation}
364 + \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
365 + \end{equation}
366 + Note that this treatment ignores rotations (and
367 + translational-rotational coupling) of the facet.  In compact systems,
368 + the facets stay relatively fixed in orientation between
369 + configurations, so this appears to be a reasonably good approximation.
370 +
371   We have implemented this method by extending the Langevin dynamics
372 < integrator in our group code, OpenMD.\cite{Meineke2005,openmd}  
372 > integrator in our code, OpenMD.\cite{Meineke2005,openmd}  At each
373 > molecular dynamics time step, the following process is carried out:
374 > \begin{enumerate}
375 > \item The standard inter-atomic forces ($\nabla_iU$) are computed.
376 > \item Delaunay triangulation is carried out using the current atomic
377 >  configuration.
378 > \item The convex hull is computed and facets are identified.
379 > \item For each facet:
380 > \begin{itemize}
381 > \item[a.] The force from the pressure bath ($-PA_f\hat{n}_f$) is
382 >  computed.
383 > \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
384 >  viscosity ($\eta$) of the bath.
385 > \item[c.] Facet drag ($-\Xi_f(t) \mathbf{v}_f(t)$) forces are
386 >  computed.
387 > \item[d.] Random forces ($\mathbf{R}_f(t)$) are computed using the
388 >  resistance tensor and the temperature ($T$) of the bath.
389 > \end{itemize}
390 > \item The facet forces are divided equally among the vertex atoms.
391 > \item Atomic positions and velocities are propagated.
392 > \end{enumerate}
393 > The Delaunay triangulation and computation of the convex hull are done
394 > using calls to the qhull library.\cite{Qhull} There is a minimal
395 > penalty for computing the convex hull and resistance tensors at each
396 > step in the molecular dynamics simulation (roughly 0.02 $\times$ cost
397 > of a single force evaluation), and the convex hull is remarkably easy
398 > to parallelize on distributed memory machines (see Appendix A).
399  
400   \section{Tests \& Applications}
401 + \label{sec:tests}
402  
403 < \subsection{Bulk modulus of gold nanoparticles}
403 > To test the new method, we have carried out simulations using the
404 > Langevin Hull on: 1) a crystalline system (gold nanoparticles), 2) a
405 > liquid droplet (SPC/E water),\cite{Berendsen1987} and 3) a
406 > heterogeneous mixture (gold nanoparticles in a water droplet). In each
407 > case, we have computed properties that depend on the external applied
408 > pressure.  Of particular interest for the single-phase systems is the
409 > isothermal compressibility,
410 > \begin{equation}
411 > \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
412 > )_{T}.
413 > \label{eq:BM}
414 > \end{equation}
415  
416 < \begin{figure}
417 < \includegraphics[width=\linewidth]{pressure_tb}
418 < \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
419 < pressure = 4 GPa}
420 < \label{pressureResponse}
421 < \end{figure}
416 > One problem with eliminating periodic boundary conditions and
417 > simulation boxes is that the volume of a three-dimensional point cloud
418 > is not well-defined.  In order to compute the compressibility of a
419 > bulk material, we make an assumption that the number density, $\rho =
420 > \frac{N}{V}$, is uniform within some region of the point cloud.  The
421 > compressibility can then be expressed in terms of the average number
422 > of particles in that region,
423 > \begin{equation}
424 > \kappa_{T} = -\frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
425 > )_{T}
426 > \label{eq:BMN}
427 > \end{equation}
428 > The region we used is a spherical volume of 10 \AA\ radius centered in
429 > the middle of the cluster. $N$ is the average number of molecules
430 > found within this region throughout a given simulation. The geometry
431 > and size of the region is arbitrary, and any bulk-like portion of the
432 > cluster can be used to compute the compressibility.
433  
434 + One might assume that the volume of the convex hull could simply be
435 + taken as the system volume $V$ in the compressibility expression
436 + (Eq. \ref{eq:BM}), but this has implications at lower pressures (which
437 + are explored in detail in the section on water droplets).
438 +
439 + The metallic force field in use for the gold nanoparticles is the
440 + quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} In all
441 + simulations involving point charges, we utilized damped shifted-force
442 + (DSF) electrostatics\cite{Fennell06} which is a variant of the Wolf
443 + summation\cite{wolf:8254} that has been shown to provide good forces
444 + and torques on molecular models for water in a computationally
445 + efficient manner.\cite{Fennell06} The damping parameter ($\alpha$) was
446 + set to 0.18 \AA$^{-1}$, and the cutoff radius was set to 12 \AA.  The
447 + Spohr potential was adopted in depicting the interaction between metal
448 + atoms and the SPC/E water molecules.\cite{ISI:000167766600035}
449 +
450 + \subsection{Compressibility of gold nanoparticles}
451 +
452 + The compressibility (and its inverse, the bulk modulus) is well-known
453 + for gold, and is captured well by the embedded atom method
454 + (EAM)~\cite{PhysRevB.33.7983} potential
455 + and related multi-body force fields.  In particular, the quantum
456 + Sutton-Chen potential gets nearly quantitative agreement with the
457 + experimental bulk modulus values, and makes a good first test of how
458 + the Langevin Hull will perform at large applied pressures.
459 +
460 + The Sutton-Chen (SC) potentials are based on a model of a metal which
461 + treats the nuclei and core electrons as pseudo-atoms embedded in the
462 + electron density due to the valence electrons on all of the other
463 + atoms in the system.\cite{Chen90} The SC potential has a simple form that closely
464 + resembles the Lennard Jones potential,
465 + \begin{equation}
466 + \label{eq:SCP1}
467 + 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] ,
468 + \end{equation}
469 + where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
470 + \begin{equation}
471 + \label{eq:SCP2}
472 + 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}}.
473 + \end{equation}
474 + $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
475 + interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
476 + Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
477 + the interactions between the valence electrons and the cores of the
478 + pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
479 + scale, $c_i$ scales the attractive portion of the potential relative
480 + to the repulsive interaction and $\alpha_{ij}$ is a length parameter
481 + that assures a dimensionless form for $\rho$. These parameters are
482 + tuned to various experimental properties such as the density, cohesive
483 + energy, and elastic moduli for FCC transition metals. The quantum
484 + Sutton-Chen (QSC) formulation matches these properties while including
485 + zero-point quantum corrections for different transition
486 + metals.\cite{PhysRevB.59.3527}
487 +
488 + In bulk gold, the experimentally-measured value for the bulk modulus
489 + is 180.32 GPa, while previous calculations on the QSC potential in
490 + periodic-boundary simulations of the bulk have yielded values of
491 + 175.53 GPa.\cite{XXX} Using the same force field, we have performed a
492 + series of relatively short (200 ps) simulations on 40 \r{A} radius
493 + nanoparticles under the Langevin Hull at a variety of applied
494 + pressures ranging from 0 GPa to XXX.  We obtain a value of 177.547 GPa
495 + for the bulk modulus for gold using this echnique.
496 +
497   \begin{figure}
498 < \includegraphics[width=\linewidth]{temperature_tb}
499 < \caption{Temperature equilibration depends on surface area and bath
500 <  viscosity.  Target Temperature = 300K}
501 < \label{temperatureResponse}
498 > \includegraphics[width=\linewidth]{stacked}
499 > \caption{The response of the internal pressure and temperature of gold
500 >  nanoparticles when first placed in the Langevin Hull
501 >  ($T_\mathrm{bath}$ = 300K, $P_\mathrm{bath}$ = 4 GPa), starting
502 >  from initial conditions that were far from the bath pressure and
503 >  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).}
504 > \label{pressureResponse}
505   \end{figure}
506  
507   \begin{equation}
# Line 250 | Line 509 | pressure = 4 GPa}
509      P}\right)
510   \end{equation}
511  
253 \begin{figure}
254 \includegraphics[width=\linewidth]{compress_tb}
255 \caption{Isothermal Compressibility (18 \AA gold nanoparticle)}
256 \label{temperatureResponse}
257 \end{figure}
258
512   \subsection{Compressibility of SPC/E water clusters}
513  
514 < 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.
514 > Prior molecular dynamics simulations on SPC/E water (both in
515 > NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
516 > ensembles) have yielded values for the isothermal compressibility that
517 > agree well with experiment.\cite{Fine1973} The results of two
518 > different approaches for computing the isothermal compressibility from
519 > Langevin Hull simulations for pressures between 1 and 6500 atm are
520 > shown in Fig. \ref{fig:compWater} along with compressibility values
521 > obtained from both other SPC/E simulations and experiment.
522 > Compressibility values from all references are for applied pressures
523 > within the range 1 - 1000 atm.
524  
525   \begin{figure}
526 < \includegraphics[width=\linewidth]{new_isothermal}
526 > \includegraphics[width=\linewidth]{new_isothermalN}
527   \caption{Compressibility of SPC/E water}
528 < \label{compWater}
528 > \label{fig:compWater}
529   \end{figure}
530  
531 < We initially used the classic compressibility formula
531 > Isothermal compressibility values calculated using the number density
532 > (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
533 > and previous simulation work throughout the 1 - 1000 atm pressure
534 > regime.  Compressibilities computed using the Hull volume, however,
535 > deviate dramatically from the experimental values at low applied
536 > pressures.  The reason for this deviation is quite simple; at low
537 > applied pressures, the liquid is in equilibrium with a vapor phase,
538 > and it is entirely possible for one (or a few) molecules to drift away
539 > from the liquid cluster (see Fig. \ref{fig:coneOfShame}).  At low
540 > pressures, the restoring forces on the facets are very gentle, and
541 > this means that the hulls often take on relatively distorted
542 > geometries which include large volumes of empty space.
543  
544 < \begin{equation}
545 < \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right )_{T}
546 < \end{equation}
544 > \begin{figure}
545 > \includegraphics[width=\linewidth]{flytest2}
546 > \caption{At low pressures, the liquid is in equilibrium with the vapor
547 >  phase, and isolated molecules can detach from the liquid droplet.
548 >  This is expected behavior, but the volume of the convex hull
549 >  includes large regions of empty space.  For this reason,
550 >  compressibilities are computed using local number densities rather
551 >  than hull volumes.}
552 > \label{fig:coneOfShame}
553 > \end{figure}
554  
555 < 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.
555 > At higher pressures, the equilibrium strongly favors the liquid phase,
556 > and the hull geometries are much more compact.  Because of the
557 > liquid-vapor effect on the convex hull, the regional number density
558 > approach (Eq. \ref{eq:BMN}) provides more reliable estimates of the
559 > compressibility.
560  
561 < 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
562 <
561 > In both the traditional compressibility formula (Eq. \ref{eq:BM}) and
562 > the number density version (Eq. \ref{eq:BMN}), multiple simulations at
563 > different pressures must be done to compute the first derivatives.  It
564 > is also possible to compute the compressibility using the fluctuation
565 > dissipation theorem using either fluctuations in the
566 > volume,\cite{Debenedetti1986},
567   \begin{equation}
568 < \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle V \right \rangle ^{2}}{V \, k_{B} \, T}
568 > \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle
569 >    V \right \rangle ^{2}}{V \, k_{B} \, T},
570   \end{equation}
571 <
572 < Thus, the compressibility of each simulation run can be calculated entirely independently from all other trajectories. However, the resulting compressibilities were still as much as an order of magnitude larger than the reference values. The effect was particularly pronounced at the low end of the pressure range. At ambient temperature and low pressures, there exists an equilibrium between vapor and liquid phases. Vapor molecules are naturally more diffuse around the exterior of the cluster, causing artificially large cluster volumes. Any compressibility calculation that relies on the hull volume will suffer these effects.
284 <
285 < In order to calculate the isothermal compressibility without being hindered by hull volume issues, we adapted the classic compressibility formula so that the compressibility could be calculated using information about the local density instead of the volume of the convex hull. We calculated the $g_{OO}(r)$ for a 1 nanosecond simulation of a cluster of 1372 SPC/E water molecules and spherically integrated the function over the bounds 0 to $r'$. In all cases, the value of $r'$ was 17.26216 $\AA$. The value of the total integral between these bounds is essentially the number (N) of molecules within volume $\frac{4}{3}\pi r'^{3}$ at a given pressure. To yield an actual molecule count, N must be scaled by an ideal density. However, even in the absence of an ideal density, we can use the relationship $\rho = \frac{N}{V}$ to rewrite the isothermal compressibility formula as
286 <
571 > or, equivalently, fluctuations in the number of molecules within the
572 > fixed region,
573   \begin{equation}
574 < \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right )_{T}
574 > \kappa_{T} = \frac{\left \langle N^{2} \right \rangle - \left \langle
575 >    N \right \rangle ^{2}}{N \, k_{B} \, T},
576   \end{equation}
577 + Thus, the compressibility of each simulation can be calculated
578 + entirely independently from all other trajectories. However, the
579 + resulting compressibilities were still as much as an order of
580 + magnitude larger than the reference values. However, compressibility
581 + calculation that relies on the hull volume will suffer these effects.
582 + WE NEED MORE HERE.
583  
291 Isothermal compressibility values calculated using this modified expression are in good agreement with the reference values throughout the 1 - 1000 atm pressure regime. Regardless of the difficulty in obtaining accurate hull volumes at low temperature and pressures, the Langevin Hull NPT method provides reasonable isothermal compressibility values for water through a large range of pressures.
292
584   \subsection{Molecular orientation distribution at cluster boundary}
585  
586 < 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.
586 > In order for non-periodic boundary conditions to be widely applicable,
587 > they must be constructed in such a way that they allow a finite system
588 > to replicate the properties of the bulk.  Early non-periodic
589 > simulation methods (e.g. hydrophobic boundary potentials) induced
590 > spurious orientational correlations deep within the simulated
591 > system.\cite{Lee1984,Belch1985} This behavior spawned many methods for
592 > fixing and characterizing the effects of artifical boundaries
593 > including methods which fix the orientations of a set of edge
594 > molecules.\cite{Warshel1978,King1989}
595  
596 < 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.
596 > As described above, the Langevin Hull does not require that the
597 > orientation of molecules be fixed, nor does it utilize an explicitly
598 > hydrophobic boundary, orientational constraint or radial constraint.
599 > Therefore, the orientational correlations of the molecules in a water
600 > cluster are of particular interest in testing this method.  Ideally,
601 > the water molecules on the surface of the cluster will have enough
602 > mobility into and out of the center of the cluster to maintain a
603 > bulk-like orientational distribution in the absence of orientational
604 > and radial constraints.  However, since the number of hydrogen bonding
605 > partners available to molecules on the exterior are limited, it is
606 > likely that there will be some effective hydrophobicity of the hull.
607  
608 < 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.
609 <
610 < 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.
611 <
303 < The orientation of a water molecule is described by
304 <
608 > To determine the extent of these effects demonstrated by the Langevin
609 > Hull, we examined the orientationations exhibited by SPC/E water in a
610 > cluster of 1372 molecules at 300 K and at pressures ranging from 1 -
611 > 1000 atm.  The orientational angle of a water molecule is described
612   \begin{equation}
613   \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
614   \end{equation}
615 + where $\vec{r}_{i}$ is the vector between molecule {\it i}'s center of
616 + mass and the cluster center of mass and $\vec{\mu}_{i}$ is the vector
617 + bisecting the H-O-H angle of molecule {\it i} Bulk-like distributions
618 + will result in $\langle \cos \theta \rangle$ values close to zero.  If
619 + the hull exhibits an overabundance of externally-oriented oxygen sites
620 + the average orientation will be negative, while dangling hydrogen
621 + sites will result in positive average orientations.
622  
623 < 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}.
624 <
623 > Fig. \ref{fig:pAngle} shows the distribution of $\cos{\theta}$ values
624 > for molecules in the interior of the cluster (squares) and for
625 > molecules included in the convex hull (circles).
626   \begin{figure}
627 < \includegraphics[width=\linewidth]{g_r_theta}
628 < \caption{Definition of coordinates}
629 < \label{coords}
627 > \includegraphics[width=\linewidth]{pAngle}
628 > \caption{Distribution of $\cos{\theta}$ values for molecules on the
629 >  interior of the cluster (squares) and for those participating in the
630 >  convex hull (circles) at a variety of pressures.  The Langevin hull
631 >  exhibits minor dewetting behavior with exposed oxygen sites on the
632 >  hull water molecules.  The orientational preference for exposed
633 >  oxygen appears to be independent of applied pressure. }
634 > \label{fig:pAngle}
635   \end{figure}
636  
637 < 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).
637 > As expected, interior molecules (those not included in the convex
638 > hull) maintain a bulk-like structure with a uniform distribution of
639 > orientations. Molecules included in the convex hull show a slight
640 > preference for values of $\cos{\theta} < 0.$ These values correspond
641 > to molecules with oxygen directed toward the exterior of the cluster,
642 > forming a dangling hydrogen bond acceptor site.
643  
644 < \begin{figure}
645 < \includegraphics[width=\linewidth]{pAngle}
646 < \caption{SPC/E water clusters: only minor dewetting at the boundary}
647 < \label{pAngle}
648 < \end{figure}
644 > In the absence of an electrostatic contribution from the exterior
645 > bath, the orientational distribution of water molecules included in
646 > the Langevin Hull will slightly resemble the distribution at a neat
647 > water liquid/vapor interface.  Previous molecular dynamics simulations
648 > of SPC/E water \cite{Taylor1996} have shown that molecules at the
649 > liquid/vapor interface favor an orientation where one hydrogen
650 > protrudes from the liquid phase. This behavior is demonstrated by
651 > experiments \cite{Du1994} \cite{Scatena2001} showing that
652 > approximately one-quarter of water molecules at the liquid/vapor
653 > interface form dangling hydrogen bonds. The negligible preference
654 > shown in these cluster simulations could be removed through the
655 > introduction of an implicit solvent model, which would provide the
656 > missing electrostatic interactions between the cluster molecules and
657 > the surrounding temperature/pressure bath.
658  
659 < 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.
659 > The orientational preference exhibited by hull molecules in the
660 > Langevin hull is significantly weaker than the preference caused by an
661 > explicit hydrophobic bounding potential.  Additionally, the Langevin
662 > Hull does not require that the orientation of any molecules be fixed
663 > in order to maintain bulk-like structure, even at the cluster surface.
664  
665 < 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.
665 > \subsection{Heterogeneous nanoparticle / water mixtures}
666  
667 < 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.
667 > \section{Discussion}
668 > \label{sec:discussion}
669  
670 + The Langevin Hull samples the isobaric-isothermal ensemble for
671 + non-periodic systems by coupling the system to an bath characterized
672 + by pressure, temperature, and solvent viscosity.  This enables the
673 + study of heterogeneous systems composed of materials of significantly
674 + different compressibilities.  Because the boundary is dynamically
675 + determined during the simulation and the molecules interacting with
676 + the boundary can change, the method and has minimal perturbations on
677 + the behavior of molecules at the edges of the simulation.  Further
678 + work on this method will involve implicit electrostatics at the
679 + boundary (which is missing in the current implementation) as well as
680 + more sophisticated treatments of the surface geometry (alpha
681 + shapes\cite{EDELSBRUNNER:1994oq,EDELSBRUNNER:1995cj} and Tight
682 + Cocone\cite{Dey:2003ts}). The non-convex hull geometries are
683 + significantly more expensive ($\mathcal{O}(N^2)$) than the convex hull
684 + ($\mathcal{O}(N \log N)$), but would enable the use of hull volumes
685 + directly in computing the compressibility of the sample.
686  
687 < \subsection{Heterogeneous nanoparticle / water mixtures}
687 > \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
688  
689 + In order to use the Langevin Hull for simulations on parallel
690 + computers, one of the more difficult tasks is to compute the bounding
691 + surface, facets, and resistance tensors when the processors have
692 + incomplete information about the entire system's topology.  Most
693 + parallel decomposition methods assign primary responsibility for the
694 + motion of an atomic site to a single processor, and we can exploit
695 + this to efficiently compute the convex hull for the entire system.
696  
697 < \section{Appendix A: Hydrodynamic tensor for triangular facets}
697 > The basic idea involves splitting the point cloud into
698 > spatially-overlapping subsets and computing the convex hulls for each
699 > of the subsets.  The points on the convex hull of the entire system
700 > are all present on at least one of the subset hulls. The algorithm
701 > works as follows:
702 > \begin{enumerate}
703 > \item Each processor computes the convex hull for its own atomic sites
704 >  (left panel in Fig. \ref{fig:parallel}).
705 > \item The Hull vertices from each processor are passed out to all of
706 >  the processors, and each processor assembles a complete list of hull
707 >  sites (this is much smaller than the original number of points in
708 >  the point cloud).
709 > \item Each processor computes the global convex hull (right panel in
710 >  Fig. \ref{fig:parallel}) using only those points that are the union
711 >  of sites gathered from all of the subset hulls.  Delaunay
712 >  triangulation is then done to obtain the facets of the global hull.
713 > \end{enumerate}
714  
715   \begin{figure}
716 < \includegraphics[width=\linewidth]{hydro}
717 < \caption{Hydro}
718 < \label{hydro}
716 > \includegraphics[width=\linewidth]{parallel}
717 > \caption{When the sites are distributed among many nodes for parallel
718 >  computation, the processors first compute the convex hulls for their
719 >  own sites (dashed lines in left panel). The positions of the sites
720 >  that make up the subset hulls are then communicated to all
721 >  processors (middle panel).  The convex hull of the system (solid line in right panel) is the convex hull of the points on the union of the subset hulls.}
722 > \label{fig:parallel}
723   \end{figure}
724  
725 < \begin{equation}
726 < \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}
727 < \end{equation}
725 > The individual hull operations scale with
726 > $\mathcal{O}(\frac{n}{p}\log\frac{n}{p})$ where $n$ is the total
727 > number of sites, and $p$ is the number of processors.  These local
728 > hull operations create a set of $p$ hulls each with approximately
729 > $\frac{n}{3pr}$ sites (for a cluster of radius $r$). The worst-case
730 > communication cost for using a ``gather'' operation to distribute this
731 > information to all processors is $\mathcal{O}( \alpha (p-1) + \frac{n
732 >  \beta (p-1)}{3 r p^2})$, while the final computation of the system
733 > hull scales as $\mathcal{O}(\frac{n}{3r}\log\frac{n}{3r})$.
734  
735 < \begin{equation}
736 < T_{if}=\frac{A_i}{8\pi\eta R_{if}}\left(I +
737 <  \frac{\mathbf{R}_{if}\mathbf{R}_{if}^T}{R_{if}^2}\right)
738 < \end{equation}
735 > For a large number of atoms on a moderately parallel machine, the
736 > total costs are dominated by the computations of the individual hulls,
737 > and communication of these hulls to so the Langevin hull sees roughly
738 > linear speed-up with increasing processor counts.
739  
740 < \section{Appendix B: Computing Convex Hulls on Parallel Computers}
353 <
354 < \section{Acknowledgments}
740 > \section*{Acknowledgments}
741   Support for this project was provided by the
742   National Science Foundation under grant CHE-0848243. Computational
743   time was provided by the Center for Research Computing (CRC) at the

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines