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 3663 by gezelter, Thu Oct 21 16:24:13 2010 UTC

# Line 80 | Line 80 | transformation to adjust the box volume.
80   motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx} the Berendsen
81   pressure bath,\cite{ISI:A1984TQ73500045} and the Langevin
82   Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize coordinate
83 < transformation to adjust the box volume.
84 <
85 < As long as the material in the simulation box is essentially a bulk
86 < liquid which has a relatively uniform compressibility, the standard
83 > transformation to adjust the box volume.  As long as the material in
84 > the simulation box is essentially a bulk-like liquid which has a
85 > relatively uniform compressibility, the standard affine transform
86   approach provides an excellent way of adjusting the volume of the
87   system and applying pressure directly via the interactions between
88 < atomic sites.  
88 > atomic sites.
89  
90 < The problem with these approaches becomes apparent when the material
90 > The problem with this approach becomes apparent when the material
91   being simulated is an inhomogeneous mixture in which portions of the
92   simulation box are incompressible relative to other portions.
93   Examples include simulations of metallic nanoparticles in liquid
# Line 97 | Line 96 | enough to avoid collisions in the incompressible regio
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.
99 > enough to avoid the instabilities in the incompressible region.
100  
101   \begin{figure}
102   \includegraphics[width=\linewidth]{AffineScale2}
# Line 111 | Line 110 | Additionally, one may often wish to simulate explicitl
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 either requires effective solute concentrations that are much
117 > higher than desirable, or unreasonable system sizes to avoid this
118 > effect.  For example, calculations using typical hydration shells
119 > solvating a protein under periodic boundary conditions are quite
120 > expensive. [CALCULATE EFFECTIVE PROTEIN CONCENTRATIONS IN TYPICAL
121 > SIMULATIONS]
122  
123 < Explicitly non-periodic systems
123 > There have been a number of other approaches to explicit
124 > non-periodicity that focus on constant or nearly-constant {\it volume}
125 > conditions while maintaining bulk-like behavior.  Berkowitz and
126 > McCammon introduced a stochastic (Langevin) boundary layer inside a
127 > region of fixed molecules which effectively enforces constant
128 > temperature and volume (NVT) conditions.\cite{Berkowitz1982} In this
129 > approach, the stochastic and fixed regions were defined relative to a
130 > central atom.  Brooks and Karplus extended this method to include
131 > deformable stochastic boundaries.\cite{iii:6312} The stochastic
132 > boundary approach has been used widely for protein
133 > simulations. [CITATIONS NEEDED]
134  
135 < Elastic Bag
135 > The electrostatic and dispersive behavior near the boundary has long
136 > been a cause for concern.  King and Warshel introduced a surface
137 > constrained all-atom solvent (SCAAS) which included polarization
138 > effects of a fixed spherical boundary to mimic bulk-like behavior
139 > without periodic boundaries.\cite{king:3647} In the SCAAS model, a
140 > layer of fixed solvent molecules surrounds the solute and any explicit
141 > solvent, and this in turn is surrounded by a continuum dielectric.
142 > MORE HERE.  WHAT DID THEY FIND?
143  
144 < Spherical Boundary approaches
144 > Beglov and Roux developed a boundary model in which the hard sphere
145 > boundary has a radius that varies with the instantaneous configuration
146 > of the solute (and solvent) molecules.\cite{beglov:9050} This model
147 > contains a clear pressure and surface tension contribution to the free
148 > energy which XXX.
149  
150 < \section{Methodology}
150 > Restraining {\it potentials} introduce repulsive potentials at the
151 > surface of a sphere or other geometry.  The solute and any explicit
152 > solvent are therefore restrained inside this potential.  Often the
153 > potentials include a weak short-range attraction to maintain the
154 > correct density at the boundary.  Beglov and Roux have also introduced
155 > a restraining boundary potential which relaxes dynamically depending
156 > on the solute geometry and the force the explicit system exerts on the
157 > shell.\cite{Beglov:1995fk}
158  
159 < We have developed a new method which uses a constant pressure and
160 < temperature bath.  This bath interacts only with the objects that are
161 < currently at the edge of the system.  Since the edge is determined
162 < dynamically as the simulation progresses, no {\it a priori} geometry
163 < is defined.  The pressure and temperature bath interacts {\it
164 <  directly} with the atoms on the edge and not with atoms interior to
165 < the simulation.  This means that there are no affine transforms
166 < required.  There are also no fictitious particles or bounding
167 < potentials used in this approach.
159 > Recently, Krilov {\it et al.} introduced a flexible boundary model
160 > that uses a Lennard-Jones potential between the solvent molecules and
161 > a boundary which is determined dynamically from the position of the
162 > nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This approach allows
163 > the confining potential to prevent solvent molecules from migrating
164 > too far from the solute surface, while providing a weak attractive
165 > force pulling the solvent molecules towards a fictitious bulk solvent.
166 > Although this approach is appealing and has physical motivation,
167 > nanoparticles do not deform far from their original geometries even at
168 > temperatures which vaporize the nearby solvent. For the systems like
169 > the one described, the flexible boundary model will be nearly
170 > identical to a fixed-volume restraining potential.
171  
172 < The basics of the method are as follows. The simulation starts as a
173 < collection of atomic locations in three dimensions (a point cloud).
174 < Delaunay triangulation is used to find all facets between coplanar
175 < neighbors.  In highly symmetric point clouds, facets can contain many
176 < atoms, but in all but the most symmetric of cases one might experience
177 < in a molecular dynamics simulation, the facets are simple triangles in
178 < 3-space that contain exactly three atoms.  
172 > The approach of Kohanoff, Caro, and Finnis is the most promising of
173 > the methods for introducing both constant pressure and temperature
174 > into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
175 > This method is based on standard Langevin dynamics, but the Brownian
176 > or random forces are allowed to act only on peripheral atoms and exert
177 > force in a direction that is inward-facing relative to the facets of a
178 > closed bounding surface.  The statistical distribution of the random
179 > forces are uniquely tied to the pressure in the external reservoir, so
180 > the method can be shown to sample the isobaric-isothermal ensemble.
181 > Kohanoff {\it et al.} used a Delaunay tessellation to generate a
182 > bounding surface surrounding the outermost atoms in the simulated
183 > system.  This is not the only possible triangulated outer surface, but
184 > guarantees that all of the random forces point inward towards the
185 > cluster.
186  
187 + In the following sections, we extend and generalize the approach of
188 + Kohanoff, Caro, and Finnis. The new method, which we are calling the
189 + ``Langevin Hull'' applies the external pressure, Langevin drag, and
190 + random forces on the facets of the {\it hull itself} instead of the
191 + atomic sites comprising the vertices of the hull.  This allows us to
192 + decouple the external pressure contribution from the drag and random
193 + force.  Section \ref{sec:meth}
194 +
195 + \section{Methodology}
196 + \label{sec:meth}
197 +
198 + We have developed a new method which uses an external bath at a fixed
199 + constant pressure ($P$) and temperature ($T$).  This bath interacts
200 + only with the objects on the exterior hull of the system.  Defining
201 + the hull of the simulation is done in a manner similar to the approach
202 + of Kohanoff, Caro and Finnis.\cite{Kohanoff:2005qm} That is, any
203 + instantaneous configuration of the atoms in the system is considered
204 + as a point cloud in three dimensional space.  Delaunay triangulation
205 + is used to find all facets between coplanar neighbors.\cite{DELAUNAY}
206 + In highly symmetric point clouds, facets can contain many atoms, but
207 + in all but the most symmetric of cases the facets are simple triangles
208 + in 3-space that contain exactly three atoms.
209 +
210   The convex hull is the set of facets that have {\it no concave
211    corners} at an atomic site.  This eliminates all facets on the
212   interior of the point cloud, leaving only those exposed to the
213   bath. Sites on the convex hull are dynamic. As molecules re-enter the
214   cluster, all interactions between atoms on that molecule and the
215 < external bath are removed.
215 > external bath are removed.  Since the edge is determined dynamically
216 > as the simulation progresses, no {\it a priori} geometry is
217 > defined. The pressure and temperature bath interacts {\it directly}
218 > with the atoms on the edge and not with atoms interior to the
219 > simulation.
220  
221 < For atomic sites in the interior of the point cloud, the equations of
222 < motion are simple Newtonian dynamics,
221 >
222 > \begin{figure}
223 > \includegraphics[width=\linewidth]{hullSample}
224 > \caption{The external temperature and pressure bath interacts only
225 >  with those atoms on the convex hull (grey surface).  The hull is
226 >  computed dynamically at each time step, and molecules dynamically
227 >  move between the interior (Newtonian) region and the Langevin hull.}
228 > \label{fig:hullSample}
229 > \end{figure}
230 >
231 >
232 > Atomic sites in the interior of the point cloud move under standard
233 > Newtonian dynamics,
234   \begin{equation}
235   m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
236   \label{eq:Newton}
# Line 164 | Line 246 | hull.  Since each vertex (or atom) provides one corner
246   \end{equation}
247  
248   The external bath interacts directly with the facets of the convex
249 < hull.  Since each vertex (or atom) provides one corner of a triangular
249 > hull. Since each vertex (or atom) provides one corner of a triangular
250   facet, the force on the facets are divided equally to each vertex.
251   However, each vertex can participate in multiple facets, so the resultant
252   force is a sum over all facets $f$ containing vertex $i$:
# Line 176 | Line 258 | coupling depends on the solvent temperature, friction
258  
259   The external pressure bath applies a force to the facets of the convex
260   hull in direct proportion to the area of the facet, while the thermal
261 < coupling depends on the solvent temperature, friction and the size and
262 < shape of each facet. The thermal interactions are expressed as a
263 < typical Langevin description of the forces,
261 > coupling depends on the solvent temperature, viscosity and the size
262 > and shape of each facet. The thermal interactions are expressed as a
263 > standard Langevin description of the forces,
264   \begin{equation}
265   \begin{array}{rclclcl}
266   {\mathbf F}_f^{\text{ext}} & = &  \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
267   & = &  -\hat{n}_f P A_f  & - & \Xi_f(t) {\mathbf v}_f(t)  & + & {\mathbf R}_f(t)
268   \end{array}
269   \end{equation}
270 < Here, $P$ is the external pressure, $A_f$ and $\hat{n}_f$ are the area
271 < and normal vectors for facet $f$, respectively.  ${\mathbf v}_f(t)$ is
272 < the velocity of the facet,
270 > Here, $A_f$ and $\hat{n}_f$ are the area and normal vectors for facet
271 > $f$, respectively.  ${\mathbf v}_f(t)$ is the velocity of the facet
272 > centroid,
273   \begin{equation}
274   {\mathbf v}_f(t) =  \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
275   \end{equation}
276 < and $\Xi_f(t)$ is a ($3 \times 3$) hydrodynamic tensor that depends on
277 < the geometry and surface area of facet $f$ and the viscosity of the
278 < fluid (See Appendix A).  The hydrodynamic tensor is related to the
276 > and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
277 > depends on the geometry and surface area of facet $f$ and the
278 > viscosity of the fluid.  The resistance tensor is related to the
279   fluctuations of the random force, $\mathbf{R}(t)$, by the
280   fluctuation-dissipation theorem,
281   \begin{eqnarray}
# Line 203 | Line 285 | Once the hydrodynamic tensor is known for a given face
285   \label{eq:randomForce}
286   \end{eqnarray}
287  
288 < Once the hydrodynamic tensor is known for a given facet (see Appendix
289 < A) obtaining a stochastic vector that has the properties in
290 < Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
291 < one-time Cholesky decomposition to obtain the square root matrix of
210 < the resistance tensor,
288 > Once the resistance tensor is known for a given facet a stochastic
289 > vector that has the properties in Eq. (\ref{eq:randomForce}) can be
290 > done efficiently by carrying out a Cholesky decomposition to obtain
291 > the square root matrix of the resistance tensor,
292   \begin{equation}
293   \Xi_f = {\bf S} {\bf S}^{T},
294   \label{eq:Cholesky}
# Line 224 | Line 305 | We have implemented this method by extending the Lange
305   random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
306   have the correct properties required by Eq. (\ref{eq:randomForce}).
307  
308 + Our treatment of the resistance tensor is approximate.  $\Xi$ for a
309 + rigid triangular plate would normally be treated as a $6 \times 6$
310 + tensor that includes translational and rotational drag as well as
311 + translational-rotational coupling. The computation of resistance
312 + tensors for rigid bodies has been detailed
313 + elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
314 + but the standard approach involving bead approximations would be
315 + prohibitively expensive if it were recomputed at each step in a
316 + molecular dynamics simulation.
317 +
318 + We are utilizing an approximate resistance tensor obtained by first
319 + constructing the Oseen tensor for the interaction of the centroid of
320 + the facet ($f$) with each of the subfacets $j$,
321 + \begin{equation}
322 + T_{jf}=\frac{A_j}{8\pi\eta R_{jf}}\left(I +
323 +  \frac{\mathbf{R}_{jf}\mathbf{R}_{jf}^T}{R_{jf}^2}\right)
324 + \end{equation}
325 + Here, $A_j$ is the area of subfacet $j$ which is a triangle containing
326 + two of the vertices of the facet along with the centroid.
327 + $\mathbf{R}_{jf}$ is the vector between the centroid of facet $f$ and
328 + the centroid of sub-facet $j$, and $I$ is the ($3 \times 3$) identity
329 + matrix.  $\eta$ is the viscosity of the external bath.
330 +
331 + \begin{figure}
332 + \includegraphics[width=\linewidth]{hydro}
333 + \caption{The resistance tensor $\Xi$ for a facet comprising sites $i$,
334 +  $j$, and $k$ is constructed using Oseen tensor contributions between
335 +  the centoid of the facet $f$ and each of the sub-facets ($i,f,j$),
336 +  ($j,f,k$), and ($k,f,i$). The centroids of the sub-facets are
337 +  located at $1$, $2$, and $3$, and the area of each sub-facet is
338 +  easily computed using half the cross product of two of the edges.}
339 + \label{hydro}
340 + \end{figure}
341 +
342 + The Oseen tensors for each of the sub-facets are added together, and
343 + the resulting matrix is inverted to give a $3 \times 3$ resistance
344 + tensor for translations of the triangular facet,
345 + \begin{equation}
346 + \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
347 + \end{equation}
348 + Note that this treatment explicitly ignores rotations (and
349 + translational-rotational coupling) of the facet.  In compact systems,
350 + the facets stay relatively fixed in orientation between
351 + configurations, so this appears to be a reasonably good approximation.
352 +
353   We have implemented this method by extending the Langevin dynamics
354 < integrator in our group code, OpenMD.\cite{Meineke2005,openmd}  
354 > integrator in our code, OpenMD.\cite{Meineke2005,openmd} The Delaunay
355 > triangulation and computation of the convex hull are done using calls
356 > to the qhull library.\cite{Qhull} There is a moderate penalty for
357 > computing the convex hull at each step in the molecular dynamics
358 > simulation (HOW MUCH?), but the convex hull is remarkably easy to
359 > parallelize on distributed memory machines (see Appendix A).
360  
361   \section{Tests \& Applications}
362 + \label{sec:tests}
363  
364 + To test the new method, we have carried out simulations using the
365 + Langevin Hull on: 1) a crystalline system (gold nanoparticles), 2) a
366 + liquid droplet (SPC/E water),\cite{SPCE} and 3) a heterogeneous
367 + mixture (gold nanoparticles in a water droplet). In each case, we have
368 + computed properties that depend on the external applied pressure.  Of
369 + particular interest for the single-phase systems is the bulk modulus,
370 + \begin{equation}
371 + \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
372 + )_{T}.
373 + \label{eq:BM}
374 + \end{equation}
375 +
376 + One problem with eliminating periodic boundary conditions and
377 + simulation boxes is that the volume of a three-dimensional point cloud
378 + is not well-defined.  In order to compute the compressibility of a
379 + bulk material, we make an assumption that the number density, $\rho =
380 + \frac{N}{V}$, is uniform within some region of the cloud.  The
381 + compressibility can then be expressed in terms of the average number
382 + of particles in that region,
383 + \begin{equation}
384 + \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
385 + )_{T}
386 + \label{eq:BMN}
387 + \end{equation}
388 + The region we used is a spherical volume of 10 \AA\ radius centered in
389 + the middle of the cluster. $N$ is the average number of molecules
390 + found within this region throughout a given simulation. The geometry
391 + and size of the region is arbitrary, and any bulk-like portion of the
392 + cluster can be used to compute the bulk modulus.
393 +
394 + One might assume that the volume of the convex hull could be taken as
395 + the system volume in the compressibility expression (Eq. \ref{eq:BM}),
396 + but this has implications at lower pressures (which are explored in
397 + detail in the section on water droplets).
398 +
399 + The metallic force field in use for the gold nanoparticles is the
400 + quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} In all
401 + simulations involving point charges, we utilized damped shifted-force
402 + (DSF) electrostatics\cite{Fennell06} which is a variant of the Wolf
403 + summation\cite{wolf:8254} that has been shown to provide good forces
404 + and torques on molecular models for water in a computationally
405 + efficient manner.\cite{Fennell06} The damping parameter ($\alpha$) was
406 + set to 0.18 \AA$^{-1}$, and the cutoff radius was set to 12 \AA.  The
407 + Spohr potential was adopted in depicting the interaction between metal
408 + atoms and the SPC/E water molecules.\cite{ISI:000167766600035}
409 +
410   \subsection{Bulk modulus of gold nanoparticles}
411  
412 + The bulk modulus is well-known for gold, and it provides a good first
413 + test of how the method compares to other similar methods.  
414 +
415 +
416   \begin{figure}
417   \includegraphics[width=\linewidth]{pressure_tb}
418   \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
# Line 258 | Line 440 | Both NVT \cite{Glattli2002} and NPT \cite{Motakabbir19
440  
441   \subsection{Compressibility of SPC/E water clusters}
442  
443 < 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.
443 > Prior molecular dynamics simulations on SPC/E water (both in
444 > NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
445 > ensembles) have yielded values for the isothermal compressibility that
446 > agree well with experiment.\cite{Fine1973} The results of two
447 > different approaches for computing the isothermal compressibility from
448 > Langevin Hull simulations for pressures between 1 and 6500 atm are
449 > shown in Fig. \ref{fig:compWater} along with compressibility values
450 > obtained from both other SPC/E simulations and experiment.
451 > Compressibility values from all references are for applied pressures
452 > within the range 1 - 1000 atm.
453  
454   \begin{figure}
455 < \includegraphics[width=\linewidth]{new_isothermal}
455 > \includegraphics[width=\linewidth]{new_isothermalN}
456   \caption{Compressibility of SPC/E water}
457 < \label{compWater}
457 > \label{fig:compWater}
458   \end{figure}
459  
460 < We initially used the classic compressibility formula
460 > Isothermal compressibility values calculated using the number density
461 > (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
462 > and previous simulation work throughout the 1 - 1000 atm pressure
463 > regime.  Compressibilities computed using the Hull volume, however,
464 > deviate dramatically from the experimental values at low applied
465 > pressures.  The reason for this deviation is quite simple; at low
466 > applied pressures, the liquid is in equilibrium with a vapor phase,
467 > and it is entirely possible for one (or a few) molecules to drift away
468 > from the liquid cluster (see Fig. \ref{fig:coneOfShame}).  At low
469 > pressures, the restoring forces on the facets are very gentle, and
470 > this means that the hulls often take on relatively distorted
471 > geometries which include large volumes of empty space.
472  
473 < \begin{equation}
474 < \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right )_{T}
475 < \end{equation}
473 > \begin{figure}
474 > \includegraphics[width=\linewidth]{flytest2}
475 > \caption{At low pressures, the liquid is in equilibrium with the vapor
476 >  phase, and isolated molecules can detach from the liquid droplet.
477 >  This is expected behavior, but the reported volume of the convex
478 >  hull includes large regions of empty space.  For this reason,
479 >  compressibilities are computed using local number densities rather
480 >  than hull volumes.}
481 > \label{fig:coneOfShame}
482 > \end{figure}
483  
484 < 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.
484 > At higher pressures, the equilibrium favors the liquid phase, and the
485 > hull geometries are much more compact.  Because of the liquid-vapor
486 > effect on the convex hull, the regional number density approach
487 > (Eq. \ref{eq:BMN}) provides more reliable estimates of the bulk
488 > modulus.
489  
490 + We initially used the classic compressibility formula to calculate the the isothermal compressibility at each target pressure. These calculations yielded compressibility values that were dramatically higher than both previous simulations and experiment. The particular compressibility expression used requires the calculation of both a volume and pressure differential, thereby stipulating that the data from at least two simulations at different pressures must be used to calculate the isothermal compressibility at one pressure.
491 +
492 + Regardless of the difficulty in obtaining accurate hull
493 + volumes at low temperature and pressures, the Langevin Hull NPT method
494 + provides reasonable isothermal compressibility values for water
495 + through a large range of pressures.
496 +
497   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
498  
499   \begin{equation}
# Line 282 | Line 502 | In order to calculate the isothermal compressibility w
502  
503   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.
504  
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
505  
287 \begin{equation}
288 \kappa_{T} = \frac{1}{N} \left ( \frac{\partial N}{\partial P} \right )_{T}
289 \end{equation}
290
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
506   \subsection{Molecular orientation distribution at cluster boundary}
507  
508   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.
# Line 328 | Line 541 | The orientational preference exhibited by hull molecul
541  
542   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.
543  
331
544   \subsection{Heterogeneous nanoparticle / water mixtures}
545  
546  
547 < \section{Appendix A: Hydrodynamic tensor for triangular facets}
547 > \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
548  
549 < \begin{figure}
338 < \includegraphics[width=\linewidth]{hydro}
339 < \caption{Hydro}
340 < \label{hydro}
341 < \end{figure}
342 <
343 < \begin{equation}
344 < \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}
345 < \end{equation}
346 <
347 < \begin{equation}
348 < T_{if}=\frac{A_i}{8\pi\eta R_{if}}\left(I +
349 <  \frac{\mathbf{R}_{if}\mathbf{R}_{if}^T}{R_{if}^2}\right)
350 < \end{equation}
351 <
352 < \section{Appendix B: Computing Convex Hulls on Parallel Computers}
353 <
354 < \section{Acknowledgments}
549 > \section*{Acknowledgments}
550   Support for this project was provided by the
551   National Science Foundation under grant CHE-0848243. Computational
552   time was provided by the Center for Research Computing (CRC) at the

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines