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 3640 by gezelter, Thu Aug 12 20:25:00 2010 UTC vs.
Revision 3665 by gezelter, Wed Oct 27 18:48:34 2010 UTC

# Line 24 | Line 24
24  
25   \title{The Langevin Hull: Constant pressure and temperature dynamics for non-periodic systems}
26  
27 < \author{Charles F. Varedeman II, Kelsey Stocker, and J. Daniel
27 > \author{Charles F. Vardeman II, Kelsey M. Stocker, and J. Daniel
28   Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
29   Department of Chemistry and Biochemistry,\\
30   University of Notre Dame\\
# 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'',
45 <  performs better than traditional affine transform methods for
46 <  systems containing heterogeneous mixtures of materials with
47 <  different compressibilities. It does not suffer from the edge
48 <  effects of boundary potential methods, and allows realistic
49 <  treatment of both external pressure and thermal conductivity to an
50 <  implicit solvents.  We apply this method to several different
51 <  systems including bare nano-particles, nano-particles in explicit
52 <  solvent, as well as clusters of liquid water and ice. The predicted
53 <  mechanical and thermal properties of these systems are in good
54 <  agreement with experimental data.
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 >  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 65 | Line 65 | Affine transform methods
65  
66   \section{Introduction}
67  
68 < Affine transform methods
68 > The most common molecular dynamics methods for sampling configurations
69 > of an isobaric-isothermal (NPT) ensemble maintain a target pressure in
70 > a simulation by coupling the volume of the system to a {\it barostat},
71 > which is an extra degree of freedom propagated along with the particle
72 > coordinates.  These methods require periodic boundary conditions,
73 > because when the instantaneous pressure in the system differs from the
74 > target pressure, the volume is reduced or expanded using {\it affine
75 >  transforms} of the system geometry. An affine transform scales the
76 > size and shape of the periodic box as well as the particle positions
77 > within the box (but not the sizes of the particles). The most common
78 > constant pressure methods, including the Melchionna
79 > modification\cite{Melchionna1993} to the Nos\'e-Hoover-Andersen
80 > equations of motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx}
81 > the Berendsen pressure bath,\cite{ISI:A1984TQ73500045} and the
82 > Langevin Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize
83 > coordinate transformation to adjust the box volume.  As long as the
84 > material in the simulation box is essentially a bulk-like liquid which
85 > has a relatively uniform compressibility, the standard affine
86 > transform approach provides an excellent way of adjusting the volume
87 > of the system and applying pressure directly via the interactions
88 > between atomic sites.
89  
90 < \begin{figure}
91 < \includegraphics[width=\linewidth]{AffineScale}
92 < \caption{Affine Scale}
93 < \label{affineScale}
94 < \end{figure}
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 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  
76
101   \begin{figure}
102   \includegraphics[width=\linewidth]{AffineScale2}
103 < \caption{Affine Scale2}
104 < \label{affineScale2}
103 > \caption{Affine Scaling constant pressure methods use box-length
104 >  scaling to adjust the volume to adjust to under- or over-pressure
105 >  conditions. In a system with a uniform compressibility (e.g. bulk
106 >  fluids) these methods can work well.  In systems containing
107 >  heterogeneous mixtures, the affine scaling moves required to adjust
108 >  the pressure in the high-compressibility regions can cause molecules
109 >  in low compressibility regions to collide.}
110 > \label{affineScale}
111   \end{figure}
112  
113 < Heterogeneous mixtures of materials with different compressibilities?
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 other approaches to explicit
125 > non-periodicity that focus on constant or nearly-constant {\it volume}
126 > conditions while maintaining bulk-like behavior.  Berkowitz and
127 > McCammon introduced a stochastic (Langevin) boundary layer inside a
128 > region of fixed molecules which effectively enforces constant
129 > temperature and volume (NVT) conditions.\cite{Berkowitz1982} In this
130 > approach, the stochastic and fixed regions were defined relative to a
131 > central atom.  Brooks and Karplus extended this method to include
132 > deformable stochastic boundaries.\cite{iii:6312} The stochastic
133 > boundary approach has been used widely for protein
134 > 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 < A new method which uses a constant pressure and temperature bath that
174 < interacts with the objects that are currently at the edge of the
175 < system.
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 < Novel features: No a priori geometry is defined, No affine transforms,
187 < No fictitious particles, No bounding potentials.
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 < Simulation starts as a collection of atomic locations in 3D (a point
203 < cloud).
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 facets of the {\it hull itself} instead of the
206 > atomic sites comprising the vertices of the hull.  This allows us to
207 > decouple the external pressure contribution from the drag and random
208 > force.  The methodology is introduced in section \ref{sec:meth}, tests
209 > on crystalline nanoparticles, liquid clusters, and heterogeneous
210 > mixtures are detailed in section \ref{sec:tests}.  Section
211 > \ref{sec:discussion} summarizes our findings.
212  
213 < Delaunay triangulation finds all facets between coplanar neighbors.
213 > \section{Methodology}
214 > \label{sec:meth}
215  
216 < The Convex Hull is the set of facets that have no concave corners at a
217 < vertex.
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 < Molecules on the convex hull are dynamic. As they re-enter the
230 < cluster, all interactions with the external bath are removed.The
231 < external bath applies pressure to the facets of the convex hull in
232 < direct proportion to the area of the facet.Thermal coupling depends on
233 < the solvent temperature, friction and the size and shape of each
234 < facet.
229 > The convex hull is the set of facets that have {\it no concave
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 + \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 dynamically
245 +  move 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
252 > m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
253 > \label{eq:Newton}
254   \end{equation}
255 <
255 > where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
256 > instantaneous velocity of site $i$ at time $t$, and $U$ is the total
257 > potential energy.  For atoms on the exterior of the cluster
258 > (i.e. those that occupy one of the vertices of the convex hull), the
259 > equation of motion is modified with an external force, ${\mathbf
260 >  F}_i^{\mathrm ext}$,
261   \begin{equation}
262 < m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}
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 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
274    F}_f^{\mathrm ext}
275   \end{equation}
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, 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 <
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 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}
137 A_f & = & \text{area of facet}\ f \\
138 \hat{n}_f & = & \text{facet normal} \\
139 P & = & \text{external pressure}
140 \end{eqnarray}
141
142 \begin{eqnarray}
143 {\mathbf v}_f(t) & = & \text{velocity of facet} \\
144 & = & \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i \\
145 \Xi_f(t) & = & \text{is a hydrodynamic tensor that depends} \\
146 & & \text{on the geometry and surface area of} \\
147 & & \text{facet} \ f\ \text{and the viscosity of the fluid.}
148 \end{eqnarray}
149
150 \begin{eqnarray}
300   \left< {\mathbf R}_f(t) \right> & = & 0 \\
301   \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
302 < \Xi_f(t)\delta(t-t^\prime)
302 > \Xi_f(t)\delta(t-t^\prime).
303 > \label{eq:randomForce}
304   \end{eqnarray}
305  
306 < Implemented in OpenMD.\cite{Meineke:2005gd,openmd}
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}
313 > \end{equation}
314 > where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
315 > vector with the statistics required for the random force can then be
316 > obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
317 > has elements chosen from a Gaussian distribution, such that:
318 > \begin{equation}
319 > \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
320 > {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
321 > \end{equation}
322 > where $\delta t$ is the timestep in use during the simulation. The
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 explicitly 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 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 done 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 + 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 + 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{Bulk modulus of gold nanoparticles}
451  
452 + The compressibility is well-known for gold, and it provides a good first
453 + test of how the method compares to other similar methods.  
454 +
455   \begin{figure}
456 < \includegraphics[width=\linewidth]{pressure_tb}
457 < \caption{Pressure response is rapid (18 \AA gold nanoparticle), target
458 < pressure = 4 GPa}
456 > \includegraphics[width=\linewidth]{P_T_combined}
457 > \caption{Pressure and temperature response of an 18 \AA\ gold
458 >  nanoparticle initially when first placed in the Langevin Hull
459 >  ($T_\mathrm{bath}$ = 300K, $P_\mathrm{bath}$ = 4 GPa) and starting
460 >  from initial conditions that were far from the bath pressure and
461 >  temperature.  The pressure response is rapid, and the thermal
462 >  equilibration depends on both total surface area and the viscosity
463 >  of the bath.}
464   \label{pressureResponse}
465   \end{figure}
466  
169 \begin{figure}
170 \includegraphics[width=\linewidth]{temperature_tb}
171 \caption{Temperature equilibration depends on surface area and bath
172  viscosity.  Target Temperature = 300K}
173 \label{temperatureResponse}
174 \end{figure}
175
467   \begin{equation}
468   \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
469      P}\right)
# Line 186 | Line 477 | pressure = 4 GPa}
477  
478   \subsection{Compressibility of SPC/E water clusters}
479  
480 + Prior molecular dynamics simulations on SPC/E water (both in
481 + NVT~\cite{Glattli2002} and NPT~\cite{Motakabbir1990, Pi2009}
482 + ensembles) have yielded values for the isothermal compressibility that
483 + agree well with experiment.\cite{Fine1973} The results of two
484 + different approaches for computing the isothermal compressibility from
485 + Langevin Hull simulations for pressures between 1 and 6500 atm are
486 + shown in Fig. \ref{fig:compWater} along with compressibility values
487 + obtained from both other SPC/E simulations and experiment.
488 + Compressibility values from all references are for applied pressures
489 + within the range 1 - 1000 atm.
490 +
491   \begin{figure}
492 < \includegraphics[width=\linewidth]{g_r_theta}
493 < \caption{Definition of coordinates}
494 < \label{coords}
492 > \includegraphics[width=\linewidth]{new_isothermalN}
493 > \caption{Compressibility of SPC/E water}
494 > \label{fig:compWater}
495   \end{figure}
496  
497 + Isothermal compressibility values calculated using the number density
498 + (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
499 + and previous simulation work throughout the 1 - 1000 atm pressure
500 + regime.  Compressibilities computed using the Hull volume, however,
501 + deviate dramatically from the experimental values at low applied
502 + pressures.  The reason for this deviation is quite simple; at low
503 + applied pressures, the liquid is in equilibrium with a vapor phase,
504 + and it is entirely possible for one (or a few) molecules to drift away
505 + from the liquid cluster (see Fig. \ref{fig:coneOfShame}).  At low
506 + pressures, the restoring forces on the facets are very gentle, and
507 + this means that the hulls often take on relatively distorted
508 + geometries which include large volumes of empty space.
509 +
510 + \begin{figure}
511 + \includegraphics[width=\linewidth]{flytest2}
512 + \caption{At low pressures, the liquid is in equilibrium with the vapor
513 +  phase, and isolated molecules can detach from the liquid droplet.
514 +  This is expected behavior, but the volume of the convex hull
515 +  includes large regions of empty space.  For this reason,
516 +  compressibilities are computed using local number densities rather
517 +  than hull volumes.}
518 + \label{fig:coneOfShame}
519 + \end{figure}
520 +
521 + At higher pressures, the equilibrium strongly favors the liquid phase,
522 + and the hull geometries are much more compact.  Because of the
523 + liquid-vapor effect on the convex hull, the regional number density
524 + approach (Eq. \ref{eq:BMN}) provides more reliable estimates of the
525 + bulk modulus.
526 +
527 + In both the traditional compressibility formula (Eq. \ref{eq:BM}) and
528 + the number density version (Eq. \ref{eq:BMN}), multiple simulations at
529 + different pressures must be done to compute the first derivatives.  It
530 + is also possible to compute the compressibility using the fluctuation
531 + dissipation theorem using either fluctuations in the
532 + volume,\cite{Debenedetti1986},
533 + \begin{equation}
534 + \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle
535 +    V \right \rangle ^{2}}{V \, k_{B} \, T},
536 + \end{equation}
537 + or, equivalently, fluctuations in the number of molecules within the
538 + fixed region,
539 + \begin{equation}
540 + \kappa_{T} = \frac{\left \langle N^{2} \right \rangle - \left \langle
541 +    N \right \rangle ^{2}}{N \, k_{B} \, T},
542 + \end{equation}
543 + Thus, the compressibility of each simulation can be calculated
544 + entirely independently from all other trajectories. However, the
545 + resulting compressibilities were still as much as an order of
546 + magnitude larger than the reference values.  Any compressibility
547 + calculation that relies on the hull volume will suffer these effects.
548 + WE NEED MORE HERE.
549 +
550 + \subsection{Molecular orientation distribution at cluster boundary}
551 +
552 + In order for non-periodic boundary conditions to be widely applicable,
553 + they must be constructed in such a way that they allow a finite system
554 + to replicate the properties of the bulk.  Naturally, this requirement
555 + has spawned many methods for fixing and characterizing the effects of
556 + artifical boundaries. Of particular interest regarding the Langevin
557 + Hull is the orientation of water molecules that are part of the
558 + geometric hull.  Ideally, all molecules in the cluster will have the
559 + same orientational distribution as bulk water.
560 +
561 + The orientation of molecules at the edges of a simulated cluster has
562 + long been a concern when performing simulations of explicitly
563 + non-periodic systems. Early work led to the surface constrained soft
564 + sphere dipole model (SCSSD) \cite{Warshel1978} in which the surface
565 + molecules are fixed in a random orientation representative of the bulk
566 + solvent structural properties. Belch, et al \cite{Belch1985} simulated
567 + clusters of TIPS2 water surrounded by a hydrophobic bounding
568 + potential. The spherical hydrophobic boundary induced dangling
569 + hydrogen bonds at the surface that propagated deep into the cluster,
570 + affecting 70\% of the 100 molecules in the simulation. This result
571 + echoes an earlier study which showed that an extended planar
572 + hydrophobic surface caused orientational preference at the surface
573 + which extended 7 \r{A} into the liquid simulation cell
574 + \cite{Lee1984}. The surface constrained all-atom solvent (SCAAS) model
575 + \cite{King1989} improved upon its SCSSD predecessor. The SCAAS model
576 + utilizes a polarization constraint which is applied to the surface
577 + molecules to maintain bulk-like structure at the cluster surface. A
578 + radial constraint is used to maintain the desired bulk density of the
579 + liquid. Both constraint forces are applied only to a pre-determined
580 + number of the outermost molecules.
581 +
582 + In contrast, the Langevin Hull does not require that the orientation
583 + of molecules be fixed, nor does it utilize an explicitly hydrophobic
584 + boundary, orientational constraint or radial constraint. The number
585 + and identity of the molecules included on the convex hull are dynamic
586 + properties, thus avoiding the formation of an artificial solvent
587 + boundary layer. The hope is that the water molecules on the surface of
588 + the cluster, if left to their own devices in the absence of
589 + orientational and radial constraints, will maintain a bulk-like
590 + orientational distribution.
591 +
592 + 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.
593 +
594 + The orientation of a water molecule is described by
595 +
596   \begin{equation}
597   \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
598   \end{equation}
599  
600 + 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}.
601 +
602   \begin{figure}
603 + \includegraphics[width=\linewidth]{g_r_theta}
604 + \caption{Definition of coordinates}
605 + \label{coords}
606 + \end{figure}
607 +
608 + 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).
609 +
610 + \begin{figure}
611   \includegraphics[width=\linewidth]{pAngle}
612   \caption{SPC/E water clusters: only minor dewetting at the boundary}
613   \label{pAngle}
614   \end{figure}
615  
616 < \begin{figure}
206 < \includegraphics[width=\linewidth]{isothermal}
207 < \caption{Compressibility of SPC/E water}
208 < \label{compWater}
209 < \end{figure}
616 > 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.
617  
618 < \subsection{Heterogeneous nanoparticle / water mixtures}
618 > 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.
619  
620 + 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.
621  
622 < \section{Appendix A: Hydrodynamic tensor for triangular facets}
622 > \subsection{Heterogeneous nanoparticle / water mixtures}
623  
624 < \begin{figure}
625 < \includegraphics[width=\linewidth]{hydro}
218 < \caption{Hydro}
219 < \label{hydro}
220 < \end{figure}
624 > \section{Discussion}
625 > \label{sec:discussion}
626  
627 < \begin{equation}
223 < \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}
224 < \end{equation}
627 > \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
628  
629 < \begin{equation}
227 < T_{if}=\frac{A_i}{8\pi\eta R_{if}}\left(I +
228 <  \frac{\mathbf{R}_{if}\mathbf{R}_{if}^T}{R_{if}^2}\right)
229 < \end{equation}
230 <
231 < \section{Appendix B: Computing Convex Hulls on Parallel Computers}
232 <
233 < \section{Acknowledgments}
629 > \section*{Acknowledgments}
630   Support for this project was provided by the
631   National Science Foundation under grant CHE-0848243. Computational
632   time was provided by the Center for Research Computing (CRC) at the

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines