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 3663 by gezelter, Thu Oct 21 16:24:13 2010 UTC vs.
Revision 3665 by gezelter, Wed Oct 27 18:48:34 2010 UTC

# 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
77 < particles). The most common constant pressure methods, including the
78 < Melchionna modification\cite{Melchionna1993} to the
79 < Nos\'e-Hoover-Andersen equations of
80 < motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx} the Berendsen
81 < pressure bath,\cite{ISI:A1984TQ73500045} and the Langevin
82 < Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize coordinate
83 < transformation to adjust the box volume.  As long as the material in
84 < the simulation box is essentially a bulk-like liquid which has a
85 < relatively uniform compressibility, the standard affine transform
86 < approach provides an excellent way of adjusting the volume of the
87 < system and applying pressure directly via the interactions between
88 < atomic sites.
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 < The problem with this approach becomes apparent when the material
91 < being simulated is an inhomogeneous mixture in which portions of the
92 < simulation box are incompressible relative to other portions.
93 < Examples include simulations of metallic nanoparticles in liquid
94 < environments, proteins at interfaces, as well as other multi-phase or
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 the instabilities 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}
# Line 113 | Line 113 | volume either requires effective solute concentrations
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
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
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 + \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
# Line 133 | Line 134 | been a cause for concern.  King and Warshel introduced
134   simulations. [CITATIONS NEEDED]
135  
136   The electrostatic and dispersive behavior near the boundary has long
137 < been a cause for concern.  King and Warshel introduced a surface
138 < constrained all-atom solvent (SCAAS) which included polarization
139 < effects of a fixed spherical boundary to mimic bulk-like behavior
140 < without periodic boundaries.\cite{king:3647} In the SCAAS model, a
141 < layer of fixed solvent molecules surrounds the solute and any explicit
142 < solvent, and this in turn is surrounded by a continuum dielectric.
143 < MORE HERE.  WHAT DID THEY FIND?
144 <
145 < Beglov and Roux developed a boundary model in which the hard sphere
146 < boundary has a radius that varies with the instantaneous configuration
147 < of the solute (and solvent) molecules.\cite{beglov:9050} This model
148 < contains a clear pressure and surface tension contribution to the free
149 < energy which XXX.
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 + 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 + \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 this potential.  Often the
167 < potentials include a weak short-range attraction to maintain the
168 < correct density at the boundary.  Beglov and Roux have also introduced
169 < a restraining boundary potential which relaxes dynamically depending
170 < on the solute geometry and the force the explicit system exerts on the
171 < shell.\cite{Beglov:1995fk}
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 < Recently, Krilov {\it et al.} introduced a flexible boundary model
174 < that uses a Lennard-Jones potential between the solvent molecules and
175 < a boundary which is determined dynamically from the position of the
176 < nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This approach allows
177 < the confining potential to prevent solvent molecules from migrating
178 < too far from the solute surface, while providing a weak attractive
179 < force pulling the solvent molecules towards a fictitious bulk solvent.
180 < Although this approach is appealing and has physical motivation,
181 < nanoparticles do not deform far from their original geometries even at
182 < temperatures which vaporize the nearby solvent. For the systems like
183 < the one described, the flexible boundary model will be nearly
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 + \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}
# Line 190 | Line 205 | force.  Section \ref{sec:meth}
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.  Section \ref{sec:meth}
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   \section{Methodology}
214   \label{sec:meth}
215  
216 < We have developed a new method which uses an external bath at a fixed
217 < constant pressure ($P$) and temperature ($T$).  This bath interacts
218 < only with the objects on the exterior hull of the system.  Defining
219 < the hull of the simulation is done in a manner similar to the approach
220 < of Kohanoff, Caro and Finnis.\cite{Kohanoff:2005qm} That is, any
221 < instantaneous configuration of the atoms in the system is considered
222 < as a point cloud in three dimensional space.  Delaunay triangulation
223 < is used to find all facets between coplanar neighbors.\cite{DELAUNAY}
224 < In highly symmetric point clouds, facets can contain many atoms, but
225 < in all but the most symmetric of cases the facets are simple triangles
226 < in 3-space that contain exactly three atoms.
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.  Since the edge is determined dynamically
235 < as the simulation progresses, no {\it a priori} geometry is
236 < defined. The pressure and temperature bath interacts {\it directly}
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  
221
240   \begin{figure}
241   \includegraphics[width=\linewidth]{hullSample}
242   \caption{The external temperature and pressure bath interacts only
# Line 228 | Line 246 | simulation.
246   \label{fig:hullSample}
247   \end{figure}
248  
249 <
232 < Atomic sites in the interior of the point cloud move under standard
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,
# Line 245 | 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 267 | Line 285 | Here, $A_f$ and $\hat{n}_f$ are the area and normal ve
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, $A_f$ and $\hat{n}_f$ are the area and normal vectors for facet
289 < $f$, respectively.  ${\mathbf v}_f(t)$ is the velocity of the facet
290 < centroid,
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}
# Line 285 | Line 303 | Once the resistance tensor is known for a given facet
303   \label{eq:randomForce}
304   \end{eqnarray}
305  
306 < Once the resistance tensor is known for a given facet a stochastic
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 < done efficiently by carrying out a Cholesky decomposition to obtain
309 < the square root matrix of the resistance tensor,
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 315 | Line 333 | We are utilizing an approximate resistance tensor obta
333   prohibitively expensive if it were recomputed at each step in a
334   molecular dynamics simulation.
335  
336 < We are utilizing an approximate resistance tensor obtained by first
337 < constructing the Oseen tensor for the interaction of the centroid of
338 < the facet ($f$) with each of the subfacets $j$,
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_{jf}=\frac{A_j}{8\pi\eta R_{jf}}\left(I +
341 <  \frac{\mathbf{R}_{jf}\mathbf{R}_{jf}^T}{R_{jf}^2}\right)
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, $A_j$ is the area of subfacet $j$ which is a triangle containing
344 < two of the vertices of the facet along with the centroid.
345 < $\mathbf{R}_{jf}$ is the vector between the centroid of facet $f$ and
346 < the centroid of sub-facet $j$, and $I$ is the ($3 \times 3$) identity
347 < matrix.  $\eta$ is the viscosity of the external bath.
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}
# Line 339 | Line 357 | The Oseen tensors for each of the sub-facets are added
357   \label{hydro}
358   \end{figure}
359  
360 < The Oseen tensors for each of the sub-facets are added together, and
361 < the resulting matrix is inverted to give a $3 \times 3$ resistance
362 < tensor for translations of the triangular facet,
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}
# Line 351 | Line 369 | integrator in our code, OpenMD.\cite{Meineke2005,openm
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} The Delaunay
373 < triangulation and computation of the convex hull are done using calls
374 < to the qhull library.\cite{Qhull} There is a moderate penalty for
375 < computing the convex hull at each step in the molecular dynamics
376 < simulation (HOW MUCH?), but the convex hull is remarkably easy to
377 < parallelize on distributed memory machines (see Appendix A).
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{SPCE} and 3) a heterogeneous
406 < mixture (gold nanoparticles in a water droplet). In each case, we have
407 < computed properties that depend on the external applied pressure.  Of
408 < particular interest for the single-phase systems is the bulk modulus,
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}.
# Line 377 | Line 417 | bulk material, we make an assumption that the number d
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 cloud.  The
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
424 > \kappa_{T} = -\frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
425   )_{T}
426   \label{eq:BMN}
427   \end{equation}
# Line 389 | Line 429 | cluster can be used to compute the bulk modulus.
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 bulk modulus.
432 > cluster can be used to compute the compressibility.
433  
434 < One might assume that the volume of the convex hull could be taken as
435 < the system volume in the compressibility expression (Eq. \ref{eq:BM}),
436 < but this has implications at lower pressures (which are explored in
437 < detail in the section on water droplets).
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
# Line 409 | Line 449 | The bulk modulus is well-known for gold, and it provid
449  
450   \subsection{Bulk modulus of gold nanoparticles}
451  
452 < The bulk modulus is well-known for gold, and it provides a good first
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  
415
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}
421 \end{figure}
422
423 \begin{figure}
424 \includegraphics[width=\linewidth]{temperature_tb}
425 \caption{Temperature equilibration depends on surface area and bath
426  viscosity.  Target Temperature = 300K}
427 \label{temperatureResponse}
465   \end{figure}
466  
467   \begin{equation}
# Line 474 | Line 511 | geometries which include large volumes of empty space.
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 reported volume of the convex
515 <  hull includes large regions of empty space.  For this reason,
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 favors the liquid phase, and the
522 < hull geometries are much more compact.  Because of the liquid-vapor
523 < effect on the convex hull, the regional number density approach
524 < (Eq. \ref{eq:BMN}) provides more reliable estimates of the bulk
525 < modulus.
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 < 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.
528 <
529 < Regardless of the difficulty in obtaining accurate hull
530 < volumes at low temperature and pressures, the Langevin Hull NPT method
531 < provides reasonable isothermal compressibility values for water
532 < 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 <
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 V \right \rangle ^{2}}{V \, k_{B} \, T}
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  
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
505
550   \subsection{Molecular orientation distribution at cluster boundary}
551  
552 < 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.
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 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.
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 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.
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  
# Line 543 | Line 621 | The orientational preference exhibited by hull molecul
621  
622   \subsection{Heterogeneous nanoparticle / water mixtures}
623  
624 + \section{Discussion}
625 + \label{sec:discussion}
626  
627   \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
628  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines