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 3666 by gezelter, Thu Oct 28 20:30:16 2010 UTC vs.
Revision 3715 by kstocke1, Mon Jan 17 21:50:02 2011 UTC

# Line 18 | Line 18
18   \setlength{\belowcaptionskip}{30 pt}
19  
20   \bibpunct{[}{]}{,}{s}{}{;}
21 < \bibliographystyle{aip}
21 > \bibliographystyle{achemso}
22  
23   \begin{document}
24  
# Line 40 | Line 40 | Notre Dame, Indiana 46556}
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 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.
43 >  to the facets to mimic contact with an external heat bath. This new
44 >  method, the ``Langevin Hull'', can handle heterogeneous mixtures of
45 >  materials with different compressibilities.  These are systems that
46 >  are problematic for traditional affine transform methods.  The
47 >  Langevin Hull does not suffer from the edge effects of boundary
48 >  potential methods, and allows realistic treatment of both external
49 >  pressure and thermal conductivity due to the presence of an implicit
50 >  solvent.  We apply this method to several different systems
51 >  including bare metal nanoparticles, nanoparticles in an explicit
52 >  solvent, as well as clusters of liquid water. The predicted
53 >  mechanical properties of these systems are in good agreement with
54 >  experimental data and previous simulation work.
55   \end{abstract}
56  
57   \newpage
# Line 66 | Line 66 | of an isobaric-isothermal (NPT) ensemble maintain a ta
66   \section{Introduction}
67  
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
69 > from an isobaric-isothermal (NPT) ensemble maintain a target pressure
70 > in a simulation by coupling the volume of the system to a {\it
71 >  barostat}, which is an extra degree of freedom propagated along with
72 > the particle coordinates.  These methods require periodic boundary
73 > conditions, because when the instantaneous pressure in the system
74 > differs from the target pressure, the volume is reduced or expanded
75 > using {\it affine transforms} of the system geometry. An affine
76 > transform scales the size and shape of the periodic box as well as the
77 > particle positions within the box (but not the sizes of the
78 > particles). The most common constant pressure methods, including the
79 > Melchionna modification\cite{Melchionna1993} to the
80 > Nos\'e-Hoover-Andersen equations of
81 > motion,\cite{Hoover85,ANDERSEN:1980vn,Sturgeon:2000kx} the Berendsen
82 > pressure bath,\cite{ISI:A1984TQ73500045} and the Langevin
83 > Piston,\cite{FELLER:1995fk,Jakobsen:2005uq} all utilize scaled
84   coordinate transformation to adjust the box volume.  As long as the
85 < material in the simulation box is essentially a bulk-like liquid which
86 < has a relatively uniform compressibility, the standard affine
87 < transform approach provides an excellent way of adjusting the volume
88 < of the system and applying pressure directly via the interactions
88 < between atomic sites.
85 > material in the simulation box has a relatively uniform
86 > compressibility, the standard affine transform approach provides an
87 > excellent way of adjusting the volume of the system and applying
88 > pressure directly via the interactions between atomic sites.
89  
90   One problem with this approach appears when the system being simulated
91   is an inhomogeneous mixture in which portions of the simulation box
# Line 100 | Line 100 | slow enough to avoid the instabilities in the incompre
100  
101   \begin{figure}
102   \includegraphics[width=\linewidth]{AffineScale2}
103 < \caption{Affine Scaling constant pressure methods use box-length
104 <  scaling to adjust the volume to adjust to under- or over-pressure
105 <  conditions. In a system with a uniform compressibility (e.g. bulk
106 <  fluids) these methods can work well.  In systems containing
107 <  heterogeneous mixtures, the affine scaling moves required to adjust
108 <  the pressure in the high-compressibility regions can cause molecules
109 <  in low compressibility regions to collide.}
103 > \caption{Affine scaling methods use box-length scaling to adjust the
104 >  volume to adjust to under- or over-pressure conditions. In a system
105 >  with a uniform compressibility (e.g. bulk fluids) these methods can
106 >  work well.  In systems containing heterogeneous mixtures, the affine
107 >  scaling moves required to adjust the pressure in the
108 >  high-compressibility regions can cause molecules in low
109 >  compressibility regions to collide.}
110   \label{affineScale}
111   \end{figure}
112  
# Line 115 | Line 115 | effect.  For example, calculations using typical hydra
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
118 > effect.  For example, calculations using typical hydration boxes
119   solvating a protein under periodic boundary conditions are quite
120 < expensive. [CALCULATE EFFECTIVE PROTEIN CONCENTRATIONS IN TYPICAL
121 < SIMULATIONS]
120 > expensive.  A 62 \AA$^3$ box of water solvating a moderately small
121 > protein like hen egg white lysozyme (PDB code: 1LYZ) yields an
122 > effective protein concentration of 100 mg/mL.\cite{Asthagiri20053300}
123  
124 + {\it Total} protein concentrations in the cell are typically on the
125 + order of 160-310 mg/ml,\cite{Brown1991195} and individual proteins
126 + have concentrations orders of magnitude lower than this in the
127 + cellular environment. The effective concentrations of single proteins
128 + in simulations may have significant effects on the structure and
129 + dynamics of simulated structures.
130 +
131   \subsection*{Boundary Methods}
132 < There have been a number of other approaches to explicit
133 < non-periodicity that focus on constant or nearly-constant {\it volume}
134 < conditions while maintaining bulk-like behavior.  Berkowitz and
135 < McCammon introduced a stochastic (Langevin) boundary layer inside a
136 < region of fixed molecules which effectively enforces constant
137 < temperature and volume (NVT) conditions.\cite{Berkowitz1982} In this
138 < approach, the stochastic and fixed regions were defined relative to a
139 < central atom.  Brooks and Karplus extended this method to include
140 < deformable stochastic boundaries.\cite{iii:6312} The stochastic
141 < boundary approach has been used widely for protein
142 < simulations. [CITATIONS NEEDED]
132 > There have been a number of approaches to handle simulations of
133 > explicitly non-periodic systems that focus on constant or
134 > nearly-constant {\it volume} conditions while maintaining bulk-like
135 > behavior.  Berkowitz and McCammon introduced a stochastic (Langevin)
136 > boundary layer inside a region of fixed molecules which effectively
137 > enforces constant temperature and volume (NVT)
138 > conditions.\cite{Berkowitz1982} In this approach, the stochastic and
139 > fixed regions were defined relative to a central atom.  Brooks and
140 > Karplus extended this method to include deformable stochastic
141 > boundaries.\cite{iii:6312} The stochastic boundary approach has been
142 > used widely for protein simulations.
143  
144   The electrostatic and dispersive behavior near the boundary has long
145   been a cause for concern when performing simulations of explicitly
# Line 142 | Line 150 | affecting most of molecules in the simulation.  This r
150   simulated clusters of TIPS2 water surrounded by a hydrophobic bounding
151   potential. The spherical hydrophobic boundary induced dangling
152   hydrogen bonds at the surface that propagated deep into the cluster,
153 < affecting most of molecules in the simulation.  This result echoes an
154 < earlier study which showed that an extended planar hydrophobic surface
155 < caused orientational preference at the surface which extended
156 < relatively deep (7 \r{A}) into the liquid simulation
157 < cell.\cite{Lee1984} The surface constrained all-atom solvent (SCAAS)
158 < model \cite{King1989} improved upon its SCSSD predecessor. The SCAAS
159 < model utilizes a polarization constraint which is applied to the
160 < surface molecules to maintain bulk-like structure at the cluster
161 < surface. A radial constraint is used to maintain the desired bulk
162 < density of the liquid. Both constraint forces are applied only to a
163 < pre-determined number of the outermost molecules.
153 > affecting most of the molecules in the simulation.  This result echoes
154 > an earlier study which showed that an extended planar hydrophobic
155 > surface caused orientational preferences at the surface which extended
156 > relatively deep (7 \AA) into the liquid simulation cell.\cite{Lee1984}
157 > The surface constrained all-atom solvent (SCAAS) model \cite{King1989}
158 > improved upon its SCSSD predecessor. The SCAAS model utilizes a
159 > polarization constraint which is applied to the surface molecules to
160 > maintain bulk-like structure at the cluster surface. A radial
161 > constraint is used to maintain the desired bulk density of the
162 > liquid. Both constraint forces are applied only to a pre-determined
163 > number of the outermost molecules.
164  
165   Beglov and Roux have developed a boundary model in which the hard
166   sphere boundary has a radius that varies with the instantaneous
167   configuration of the solute (and solvent) molecules.\cite{beglov:9050}
168   This model contains a clear pressure and surface tension contribution
169 < to the free energy which XXX.
169 > to the free energy.
170  
171   \subsection*{Restraining Potentials}
172   Restraining {\it potentials} introduce repulsive potentials at the
# Line 173 | Line 181 | position of the nearest solute atom.\cite{LiY._jp04685
181   Recently, Krilov {\it et al.} introduced a {\it flexible} boundary
182   model that uses a Lennard-Jones potential between the solvent
183   molecules and a boundary which is determined dynamically from the
184 < position of the nearest solute atom.\cite{LiY._jp046852t,Zhu:xw} This
184 > position of the nearest solute atom.\cite{LiY._jp046852t,Zhu:2008fk} This
185   approach allows the confining potential to prevent solvent molecules
186   from migrating too far from the solute surface, while providing a weak
187   attractive force pulling the solvent molecules towards a fictitious
# Line 189 | Line 197 | force in a direction that is inward-facing relative to
197   into non-periodic simulations.\cite{Kohanoff:2005qm,Baltazar:2006ru}
198   This method is based on standard Langevin dynamics, but the Brownian
199   or random forces are allowed to act only on peripheral atoms and exert
200 < force in a direction that is inward-facing relative to the facets of a
201 < closed bounding surface.  The statistical distribution of the random
200 > forces in a direction that is inward-facing relative to the facets of
201 > a closed bounding surface.  The statistical distribution of the random
202   forces are uniquely tied to the pressure in the external reservoir, so
203   the method can be shown to sample the isobaric-isothermal ensemble.
204   Kohanoff {\it et al.} used a Delaunay tessellation to generate a
# Line 202 | Line 210 | random forces on the facets of the {\it hull itself} i
210   In the following sections, we extend and generalize the approach of
211   Kohanoff, Caro, and Finnis. The new method, which we are calling the
212   ``Langevin Hull'' applies the external pressure, Langevin drag, and
213 < random forces on the facets of the {\it hull itself} instead of the
214 < atomic sites comprising the vertices of the hull.  This allows us to
215 < decouple the external pressure contribution from the drag and random
216 < force.  The methodology is introduced in section \ref{sec:meth}, tests
217 < on crystalline nanoparticles, liquid clusters, and heterogeneous
218 < mixtures are detailed in section \ref{sec:tests}.  Section
219 < \ref{sec:discussion} summarizes our findings.
213 > random forces on the {\it facets of the hull} instead of the atomic
214 > sites comprising the vertices of the hull.  This allows us to decouple
215 > the external pressure contribution from the drag and random force.
216 > The methodology is introduced in section \ref{sec:meth}, tests on
217 > crystalline nanoparticles, liquid clusters, and heterogeneous mixtures
218 > are detailed in section \ref{sec:tests}.  Section \ref{sec:discussion}
219 > summarizes our findings.
220  
221   \section{Methodology}
222   \label{sec:meth}
223  
224   The Langevin Hull uses an external bath at a fixed constant pressure
225 < ($P$) and temperature ($T$).  This bath interacts only with the
226 < objects on the exterior hull of the system.  Defining the hull of the
227 < simulation is done in a manner similar to the approach of Kohanoff,
228 < Caro and Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous
229 < configuration of the atoms in the system is considered as a point
230 < cloud in three dimensional space.  Delaunay triangulation is used to
231 < find all facets between coplanar
232 < neighbors.\cite{delaunay,springerlink:10.1007/BF00977785}  In highly
225 > ($P$) and temperature ($T$) with an effective solvent viscosity
226 > ($\eta$).  This bath interacts only with the objects on the exterior
227 > hull of the system.  Defining the hull of the atoms in a simulation is
228 > done in a manner similar to the approach of Kohanoff, Caro and
229 > Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous configuration
230 > of the atoms in the system is considered as a point cloud in three
231 > dimensional space.  Delaunay triangulation is used to find all facets
232 > between coplanar
233 > neighbors.\cite{delaunay,springerlink:10.1007/BF00977785} In highly
234   symmetric point clouds, facets can contain many atoms, but in all but
235 < the most symmetric of cases the facets are simple triangles in 3-space
236 < that contain exactly three atoms.
235 > the most symmetric of cases, the facets are simple triangles in
236 > 3-space which contain exactly three atoms.
237  
238   The convex hull is the set of facets that have {\it no concave
239    corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
# Line 238 | Line 247 | simulation.
247   simulation.
248  
249   \begin{figure}
250 < \includegraphics[width=\linewidth]{hullSample}
250 > \includegraphics[width=\linewidth]{solvatedNano}
251   \caption{The external temperature and pressure bath interacts only
252    with those atoms on the convex hull (grey surface).  The hull is
253 <  computed dynamically at each time step, and molecules dynamically
254 <  move between the interior (Newtonian) region and the Langevin hull.}
253 >  computed dynamically at each time step, and molecules can move
254 >  between the interior (Newtonian) region and the Langevin Hull.}
255   \label{fig:hullSample}
256   \end{figure}
257  
# Line 257 | Line 266 | equation of motion is modified with an external force,
266   potential energy.  For atoms on the exterior of the cluster
267   (i.e. those that occupy one of the vertices of the convex hull), the
268   equation of motion is modified with an external force, ${\mathbf
269 <  F}_i^{\mathrm ext}$,
269 >  F}_i^{\mathrm ext}$:
270   \begin{equation}
271   m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
272   \end{equation}
# Line 293 | Line 302 | viscosity of the fluid.  The resistance tensor is rela
302   \end{equation}
303   and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
304   depends on the geometry and surface area of facet $f$ and the
305 < viscosity of the fluid.  The resistance tensor is related to the
305 > viscosity of the bath.  The resistance tensor is related to the
306   fluctuations of the random force, $\mathbf{R}(t)$, by the
307   fluctuation-dissipation theorem,
308   \begin{eqnarray}
# Line 323 | Line 332 | Our treatment of the resistance tensor is approximate.
332   random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
333   have the correct properties required by Eq. (\ref{eq:randomForce}).
334  
335 < Our treatment of the resistance tensor is approximate.  $\Xi$ for a
335 > Our treatment of the resistance tensor is approximate.  $\Xi_f$ for a
336   rigid triangular plate would normally be treated as a $6 \times 6$
337   tensor that includes translational and rotational drag as well as
338   translational-rotational coupling. The computation of resistance
# Line 363 | Line 372 | Note that this treatment explicitly ignores rotations
372   \begin{equation}
373   \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
374   \end{equation}
375 < Note that this treatment explicitly ignores rotations (and
375 > Note that this treatment ignores rotations (and
376   translational-rotational coupling) of the facet.  In compact systems,
377   the facets stay relatively fixed in orientation between
378   configurations, so this appears to be a reasonably good approximation.
# Line 373 | Line 382 | molecular dynamics time step, the following process is
382   molecular dynamics time step, the following process is carried out:
383   \begin{enumerate}
384   \item The standard inter-atomic forces ($\nabla_iU$) are computed.
385 < \item Delaunay triangulation is done using the current atomic
385 > \item Delaunay triangulation is carried out using the current atomic
386    configuration.
387   \item The convex hull is computed and facets are identified.
388   \item For each facet:
389   \begin{itemize}
390 < \item[a.] The force from the pressure bath ($-PA_f\hat{n}_f$) is
390 > \item[a.] The force from the pressure bath ($-\hat{n}_fPA_f$) is
391    computed.
392   \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
393    viscosity ($\eta$) of the bath.
# Line 403 | Line 412 | heterogeneous mixture (gold nanoparticles in a water d
412   To test the new method, we have carried out simulations using the
413   Langevin Hull on: 1) a crystalline system (gold nanoparticles), 2) a
414   liquid droplet (SPC/E water),\cite{Berendsen1987} and 3) a
415 < 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,
415 > heterogeneous mixture (gold nanoparticles in an SPC/E water droplet). In each case, we have computed properties that depend on the external applied pressure. Of particular interest for the single-phase systems is the isothermal compressibility,
416   \begin{equation}
417   \kappa_{T} = -\frac{1}{V} \left ( \frac{\partial V}{\partial P} \right
418   )_{T}.
# Line 415 | Line 421 | is not well-defined.  In order to compute the compress
421  
422   One problem with eliminating periodic boundary conditions and
423   simulation boxes is that the volume of a three-dimensional point cloud
424 < is not well-defined.  In order to compute the compressibility of a
424 > is not well-defined. In order to compute the compressibility of a
425   bulk material, we make an assumption that the number density, $\rho =
426 < \frac{N}{V}$, is uniform within some region of the point cloud.  The
426 > \frac{N}{V}$, is uniform within some region of the point cloud. The
427   compressibility can then be expressed in terms of the average number
428   of particles in that region,
429   \begin{equation}
430   \kappa_{T} = -\frac{1}{N} \left ( \frac{\partial N}{\partial P} \right
431 < )_{T}
431 > )_{T}.
432   \label{eq:BMN}
433   \end{equation}
434 < The region we used is a spherical volume of 10 \AA\ radius centered in
434 > The region we used is a spherical volume of 20 \AA\ radius centered in
435   the middle of the cluster. $N$ is the average number of molecules
436   found within this region throughout a given simulation. The geometry
437   and size of the region is arbitrary, and any bulk-like portion of the
# Line 447 | Line 453 | atoms and the SPC/E water molecules.\cite{ISI:00016776
453   Spohr potential was adopted in depicting the interaction between metal
454   atoms and the SPC/E water molecules.\cite{ISI:000167766600035}
455  
456 < \subsection{Bulk modulus of gold nanoparticles}
456 > \subsection{Bulk Modulus of gold nanoparticles}
457  
458 < The compressibility is well-known for gold, and it provides a good first
459 < test of how the method compares to other similar methods.  
460 <
461 < \begin{figure}
462 < \includegraphics[width=\linewidth]{P_T_combined}
463 < \caption{Pressure and temperature response of an 18 \AA\ gold
464 <  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}
458 > The compressibility (and its inverse, the bulk modulus) is well-known
459 > for gold, and is captured well by the embedded atom method
460 > (EAM)~\cite{PhysRevB.33.7983} potential and related multi-body force
461 > fields.  In particular, the quantum Sutton-Chen potential gets nearly
462 > quantitative agreement with the experimental bulk modulus values, and
463 > makes a good first test of how the Langevin Hull will perform at large
464 > applied pressures.
465  
466 + The Sutton-Chen (SC) potentials are based on a model of a metal which
467 + treats the nuclei and core electrons as pseudo-atoms embedded in the
468 + electron density due to the valence electrons on all of the other
469 + atoms in the system.\cite{Chen90} The SC potential has a simple form
470 + that closely resembles the Lennard Jones potential,
471   \begin{equation}
472 < \kappa_T=-\frac{1}{V_{\mathrm{eq}}}\left(\frac{\partial V}{\partial
473 <    P}\right)
472 > \label{eq:SCP1}
473 > U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
474   \end{equation}
475 + where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
476 + \begin{equation}
477 + \label{eq:SCP2}
478 + V^{pair}_{ij}(r)=\left( \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}.
479 + \end{equation}
480 + $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
481 + interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
482 + Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
483 + the interactions between the valence electrons and the cores of the
484 + pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
485 + scale, $c_i$ scales the attractive portion of the potential relative
486 + to the repulsive interaction and $\alpha_{ij}$ is a length parameter
487 + that assures a dimensionless form for $\rho$. These parameters are
488 + tuned to various experimental properties such as the density, cohesive
489 + energy, and elastic moduli for FCC transition metals. The quantum
490 + Sutton-Chen (QSC) formulation matches these properties while including
491 + zero-point quantum corrections for different transition
492 + metals.\cite{PhysRevB.59.3527,QSC}
493  
494 + In bulk gold, the experimentally-measured value for the bulk modulus
495 + is 180.32 GPa, while previous calculations on the QSC potential in
496 + periodic-boundary simulations of the bulk crystal have yielded values
497 + of 175.53 GPa.\cite{QSC} Using the same force field, we have performed
498 + a series of 1 ns simulations on 40 \AA~ radius
499 + nanoparticles under the Langevin Hull at a variety of applied
500 + pressures ranging from 0 -- 10 GPa.  We obtain a value of 177.55 GPa
501 + for the bulk modulus of gold using this technique, in close agreement
502 + with both previous simulations and the experimental bulk modulus of
503 + gold.
504 +
505   \begin{figure}
506 < \includegraphics[width=\linewidth]{compress_tb}
507 < \caption{Isothermal Compressibility (18 \AA gold nanoparticle)}
508 < \label{temperatureResponse}
506 > \includegraphics[width=\linewidth]{stacked}
507 > \caption{The response of the internal pressure and temperature of gold
508 >  nanoparticles when first placed in the Langevin Hull
509 >  ($T_\mathrm{bath}$ = 300K, $P_\mathrm{bath}$ = 4 GPa), starting
510 >  from initial conditions that were far from the bath pressure and
511 >  temperature.  The pressure response is rapid (after the breathing mode oscillations in the nanoparticle die out), and the rate of thermal equilibration depends on both exposed surface area (top panel) and the viscosity of the bath (middle panel).}
512 > \label{fig:pressureResponse}
513   \end{figure}
514  
515 + We note that the Langevin Hull produces rapidly-converging behavior
516 + for structures that are started far from equilibrium.  In
517 + Fig. \ref{fig:pressureResponse} we show how the pressure and
518 + temperature respond to the Langevin Hull for nanoparticles that were
519 + initialized far from the target pressure and temperature.  As
520 + expected, the rate at which thermal equilibrium is achieved depends on
521 + the total surface area of the cluster exposed to the bath as well as
522 + the bath viscosity.  Pressure that is applied suddenly to a cluster
523 + can excite breathing vibrations, but these rapidly damp out (on time
524 + scales of 30 -- 50 ps).
525 +
526   \subsection{Compressibility of SPC/E water clusters}
527  
528   Prior molecular dynamics simulations on SPC/E water (both in
# Line 485 | Line 533 | Compressibility values from all references are for app
533   Langevin Hull simulations for pressures between 1 and 6500 atm are
534   shown in Fig. \ref{fig:compWater} along with compressibility values
535   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.
536  
537   \begin{figure}
538   \includegraphics[width=\linewidth]{new_isothermalN}
# Line 496 | Line 542 | and previous simulation work throughout the 1 - 1000 a
542  
543   Isothermal compressibility values calculated using the number density
544   (Eq. \ref{eq:BMN}) expression are in good agreement with experimental
545 < and previous simulation work throughout the 1 - 1000 atm pressure
545 > and previous simulation work throughout the 1 -- 1000 atm pressure
546   regime.  Compressibilities computed using the Hull volume, however,
547   deviate dramatically from the experimental values at low applied
548 < pressures.  The reason for this deviation is quite simple; at low
548 > pressures.  The reason for this deviation is quite simple: at low
549   applied pressures, the liquid is in equilibrium with a vapor phase,
550   and it is entirely possible for one (or a few) molecules to drift away
551   from the liquid cluster (see Fig. \ref{fig:coneOfShame}).  At low
# Line 508 | Line 554 | geometries which include large volumes of empty space.
554   geometries which include large volumes of empty space.
555  
556   \begin{figure}
557 < \includegraphics[width=\linewidth]{flytest2}
557 > \includegraphics[width=\linewidth]{coneOfShame}
558   \caption{At low pressures, the liquid is in equilibrium with the vapor
559    phase, and isolated molecules can detach from the liquid droplet.
560    This is expected behavior, but the volume of the convex hull
561 <  includes large regions of empty space.  For this reason,
561 >  includes large regions of empty space. For this reason,
562    compressibilities are computed using local number densities rather
563    than hull volumes.}
564   \label{fig:coneOfShame}
# Line 522 | Line 568 | bulk modulus.
568   and the hull geometries are much more compact.  Because of the
569   liquid-vapor effect on the convex hull, the regional number density
570   approach (Eq. \ref{eq:BMN}) provides more reliable estimates of the
571 < bulk modulus.
571 > compressibility.
572  
573   In both the traditional compressibility formula (Eq. \ref{eq:BM}) and
574   the number density version (Eq. \ref{eq:BMN}), multiple simulations at
575   different pressures must be done to compute the first derivatives.  It
576   is also possible to compute the compressibility using the fluctuation
577   dissipation theorem using either fluctuations in the
578 < volume,\cite{Debenedetti1986},
578 > volume,\cite{Debenedetti1986}
579   \begin{equation}
580   \kappa_{T} = \frac{\left \langle V^{2} \right \rangle - \left \langle
581      V \right \rangle ^{2}}{V \, k_{B} \, T},
582 + \label{eq:BMVfluct}
583   \end{equation}
584   or, equivalently, fluctuations in the number of molecules within the
585   fixed region,
586   \begin{equation}
587   \kappa_{T} = \frac{\left \langle N^{2} \right \rangle - \left \langle
588 <    N \right \rangle ^{2}}{N \, k_{B} \, T},
588 >    N \right \rangle ^{2}}{N \, k_{B} \, T}.
589 > \label{eq:BMNfluct}
590   \end{equation}
591   Thus, the compressibility of each simulation can be calculated
592 < entirely independently from all other trajectories. However, the
593 < resulting compressibilities were still as much as an order of
594 < magnitude larger than the reference values.  Any compressibility
595 < calculation that relies on the hull volume will suffer these effects.
596 < WE NEED MORE HERE.
592 > entirely independently from other trajectories.  Compressibility
593 > calculations that rely on the hull volume will still suffer the
594 > effects of the empty space due to the vapor phase; for this reason, we
595 > recommend using the number density (Eq. \ref{eq:BMN}) or number
596 > density fluctuations (Eq. \ref{eq:BMNfluct}) for computing
597 > compressibilities.
598  
599   \subsection{Molecular orientation distribution at cluster boundary}
600  
601 < In order for non-periodic boundary conditions to be widely applicable,
602 < they must be constructed in such a way that they allow a finite system
603 < to replicate the properties of the bulk.  Naturally, this requirement
604 < has spawned many methods for fixing and characterizing the effects of
605 < artifical boundaries. Of particular interest regarding the Langevin
606 < Hull is the orientation of water molecules that are part of the
607 < geometric hull.  Ideally, all molecules in the cluster will have the
608 < same orientational distribution as bulk water.
601 > In order for a non-periodic boundary method to be widely applicable,
602 > it must be constructed in such a way that they allow a finite system
603 > to replicate the properties of the bulk. Early non-periodic simulation
604 > methods (e.g. hydrophobic boundary potentials) induced spurious
605 > orientational correlations deep within the simulated
606 > system.\cite{Lee1984,Belch1985} This behavior spawned many methods for
607 > fixing and characterizing the effects of artifical boundaries
608 > including methods which fix the orientations of a set of edge
609 > molecules.\cite{Warshel1978,King1989}
610  
611 < The orientation of molecules at the edges of a simulated cluster has
612 < long been a concern when performing simulations of explicitly
613 < non-periodic systems. Early work led to the surface constrained soft
614 < sphere dipole model (SCSSD) \cite{Warshel1978} in which the surface
615 < molecules are fixed in a random orientation representative of the bulk
616 < solvent structural properties. Belch, et al \cite{Belch1985} simulated
617 < clusters of TIPS2 water surrounded by a hydrophobic bounding
618 < potential. The spherical hydrophobic boundary induced dangling
619 < hydrogen bonds at the surface that propagated deep into the cluster,
620 < affecting 70\% of the 100 molecules in the simulation. This result
621 < 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.
611 > As described above, the Langevin Hull does not require that the
612 > orientation of molecules be fixed, nor does it utilize an explicitly
613 > hydrophobic boundary, or orientational or radial constraints.
614 > Therefore, the orientational correlations of the molecules in water
615 > clusters are of particular interest in testing this method.  Ideally,
616 > the water molecules on the surfaces of the clusters will have enough
617 > mobility into and out of the center of the cluster to maintain
618 > bulk-like orientational distribution in the absence of orientational
619 > and radial constraints.  However, since the number of hydrogen bonding
620 > partners available to molecules on the exterior are limited, it is
621 > likely that there will be an effective hydrophobicity of the hull.
622  
623 < In contrast, the Langevin Hull does not require that the orientation
624 < of molecules be fixed, nor does it utilize an explicitly hydrophobic
625 < boundary, orientational constraint or radial constraint. The number
626 < 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 <
623 > To determine the extent of these effects, we examined the
624 > orientations exhibited by SPC/E water in a cluster of 1372
625 > molecules at 300 K and at pressures ranging from 1 -- 1000 atm.  The
626 > orientational angle of a water molecule is described by
627   \begin{equation}
628   \cos{\theta}=\frac{\vec{r}_i\cdot\vec{\mu}_i}{|\vec{r}_i||\vec{\mu}_i|}
629   \end{equation}
630 + where $\vec{r}_{i}$ is the vector between molecule {\it i}'s center of
631 + mass and the cluster center of mass, and $\vec{\mu}_{i}$ is the vector
632 + bisecting the H-O-H angle of molecule {\it i}.  Bulk-like
633 + distributions will result in $\langle \cos \theta \rangle$ values
634 + close to zero.  If the hull exhibits an overabundance of
635 + externally-oriented oxygen sites, the average orientation will be
636 + negative, while dangling hydrogen sites will result in positive
637 + average orientations.
638  
639 < 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}.
640 <
639 > Fig. \ref{fig:pAngle} shows the distribution of $\cos{\theta}$ values
640 > for molecules in the interior of the cluster (squares) and for
641 > molecules included in the convex hull (circles).
642   \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}
643   \includegraphics[width=\linewidth]{pAngle}
644 < \caption{SPC/E water clusters: only minor dewetting at the boundary}
645 < \label{pAngle}
644 > \caption{Distribution of $\cos{\theta}$ values for molecules on the
645 >  interior of the cluster (squares) and for those participating in the
646 >  convex hull (circles) at a variety of pressures.  The Langevin Hull
647 >  exhibits minor dewetting behavior with exposed oxygen sites on the
648 >  hull water molecules.  The orientational preference for exposed
649 >  oxygen appears to be independent of applied pressure. }
650 > \label{fig:pAngle}
651   \end{figure}
652  
653 < 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.
653 > As expected, interior molecules (those not included in the convex
654 > hull) maintain a bulk-like structure with a uniform distribution of
655 > orientations. Molecules included in the convex hull show a slight
656 > preference for values of $\cos{\theta} < 0.$ These values correspond
657 > to molecules with oxygen directed toward the exterior of the cluster,
658 > forming dangling hydrogen bond acceptor sites.
659  
660 < 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.
660 > The orientational preference exhibited by water molecules on the hull
661 > is significantly weaker than the preference caused by an explicit
662 > hydrophobic bounding potential.  Additionally, the Langevin Hull does
663 > not require that the orientation of any molecules be fixed in order to
664 > maintain bulk-like structure, even near the cluster surface.
665  
666 < 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.
666 > Previous molecular dynamics simulations of SPC/E liquid / vapor
667 > interfaces using periodic boundary conditions have shown that
668 > molecules on the liquid side of interface favor a similar orientation
669 > where oxygen is directed away from the bulk.\cite{Taylor1996} These
670 > simulations had well-defined liquid and vapor phase regions
671 > equilibrium and it was observed that {\it vapor} molecules generally
672 > had one hydrogen protruding from the surface, forming a dangling
673 > hydrogen bond donor. Our water clusters do not have a true vapor
674 > region, but rather a few transient molecules that leave the liquid
675 > droplet (and which return to the droplet relatively quickly).
676 > Although we cannot obtain an orientational preference of vapor phase
677 > molecules in a Langevin Hull simulation, but we do agree with previous
678 > estimates of the orientation of {\it liquid phase} molecules at the
679 > interface.
680  
681   \subsection{Heterogeneous nanoparticle / water mixtures}
682  
683 + To further test the method, we simulated gold nanopartices ($r = 18$
684 + \AA) solvated by explicit SPC/E water clusters using a model for the
685 + gold / water interactions that has been used by Dou {\it et. al.} for
686 + investigating the separation of water films near hot metal
687 + surfaces.\cite{ISI:000167766600035} The Langevin Hull was used to
688 + sample pressures of 1, 2, 5, 10, 20, 50, 100 and 200 atm, while all
689 + simulations were done at a temperature of 300 K.   At these
690 + temperatures and pressures, there is no observed separation of the
691 + water film from the surface.  
692 +
693 + In Fig. \ref{fig:RhoR} we show the density of water and gold as a
694 + function of the distance from the center of the nanoparticle.  Higher
695 + applied pressures appear to destroy structural correlations in the
696 + outermost monolayer of the gold nanoparticle as well as in the water
697 + at the near the metal / water interface.  Simulations at increased
698 + pressures exhibit significant overlap of the gold and water densities,
699 + indicating a less well-defined interfacial surface.
700 +
701 + \begin{figure}
702 + \includegraphics[width=\linewidth]{RhoR}
703 + \caption{Density profiles of gold and water at the nanoparticle
704 +  surface. Each curve has been normalized by the average density in
705 +  the bulk-like region available to the corresponding material.  Higher applied pressures
706 +  de-structure both the gold nanoparticle surface and water at the
707 +  metal/water interface.}
708 + \label{fig:RhoR}
709 + \end{figure}
710 +
711 + At even higher pressures (500 atm and above), problems with the metal
712 + - water interaction potential became quite clear.  The model we are
713 + using appears to have been parameterized for relatively low pressures;
714 + it utilizes both shifted Morse and repulsive Morse potentials to model
715 + the Au/O and Au/H interactions, respectively.  The repulsive wall of
716 + the Morse potential does not diverge quickly enough at short distances
717 + to prevent water from diffusing into the center of the gold
718 + nanoparticles.  This behavior is likely not a realistic description of
719 + the real physics of the situation.  A better model of the gold-water
720 + adsorption behavior appears to require harder repulsive walls to
721 + prevent this behavior.
722 +
723   \section{Discussion}
724   \label{sec:discussion}
725  
726 + The Langevin Hull samples the isobaric-isothermal ensemble for
727 + non-periodic systems by coupling the system to a bath characterized by
728 + pressure, temperature, and solvent viscosity.  This enables the
729 + simulation of heterogeneous systems composed of materials with
730 + significantly different compressibilities.  Because the boundary is
731 + dynamically determined during the simulation and the molecules
732 + interacting with the boundary can change, the method inflicts minimal
733 + perturbations on the behavior of molecules at the edges of the
734 + simulation.  Further work on this method will involve implicit
735 + electrostatics at the boundary (which is missing in the current
736 + implementation) as well as more sophisticated treatments of the
737 + surface geometry (alpha
738 + shapes\cite{EDELSBRUNNER:1994oq,EDELSBRUNNER:1995cj} and Tight
739 + Cocone\cite{Dey:2003ts}). The non-convex hull geometries are
740 + significantly more expensive ($\mathcal{O}(N^2)$) than the convex hull
741 + ($\mathcal{O}(N \log N)$), but would enable the use of hull volumes
742 + directly in computing the compressibility of the sample.
743 +
744   \section*{Appendix A: Computing Convex Hulls on Parallel Computers}
745  
746   In order to use the Langevin Hull for simulations on parallel
747   computers, one of the more difficult tasks is to compute the bounding
748 < surface, facets, and resistance tensors when the processors have
749 < incomplete information about the entire system's topology.  Most
748 > surface, facets, and resistance tensors when the individual processors
749 > have incomplete information about the entire system's topology.  Most
750   parallel decomposition methods assign primary responsibility for the
751   motion of an atomic site to a single processor, and we can exploit
752   this to efficiently compute the convex hull for the entire system.
753  
754 < The basic idea is that if we split the original point cloud into
755 < spatially-overlapping subsets and compute the convex hulls for each of
756 < the subsets, the points on the convex hull of the entire system are
757 < all present on at least one of the subset hulls.  The algorithm works
758 < as follows:
754 > The basic idea involves splitting the point cloud into
755 > spatially-overlapping subsets and computing the convex hulls for each
756 > of the subsets.  The points on the convex hull of the entire system
757 > are all present on at least one of the subset hulls. The algorithm
758 > works as follows:
759   \begin{enumerate}
760   \item Each processor computes the convex hull for its own atomic sites
761 <  (dashed lines in Fig. \ref{fig:parallel}).
762 < \item The Hull vertices from each processor are passed out to all of
761 >  (left panel in Fig. \ref{fig:parallel}).
762 > \item The Hull vertices from each processor are communicated to all of
763    the processors, and each processor assembles a complete list of hull
764    sites (this is much smaller than the original number of points in
765    the point cloud).
766 < \item Each processor computes the convex hull of these sites (solid
767 <  line in Fig. \ref{fig:parallel}) and carries out Delaunay
768 <  triangulation to obtain the facets of the global hull.
766 > \item Each processor computes the global convex hull (right panel in
767 >  Fig. \ref{fig:parallel}) using only those points that are the union
768 >  of sites gathered from all of the subset hulls.  Delaunay
769 >  triangulation is then done to obtain the facets of the global hull.
770   \end{enumerate}
771  
772   \begin{figure}
773 < \begin{centering}
656 < \includegraphics[width=3in]{parallel}
773 > \includegraphics[width=\linewidth]{parallel}
774   \caption{When the sites are distributed among many nodes for parallel
775    computation, the processors first compute the convex hulls for their
776 <  own sites (dashed lines in upper panel). The positions of the sites
777 <  that make up the convex hulls are then communicated to all
778 <  processors.  The convex hull of the system (solid line in lower
779 <  panel) is the convex hull of the points on the hulls for all
780 <  processors. }
664 < \end{centering}
776 >  own sites (dashed lines in left panel). The positions of the sites
777 >  that make up the subset hulls are then communicated to all
778 >  processors (middle panel).  The convex hull of the system (solid line in
779 >  right panel) is the convex hull of the points on the union of the subset
780 >  hulls.}
781   \label{fig:parallel}
782   \end{figure}
783  
784   The individual hull operations scale with
785 < $O(\frac{n}{p}\log\frac{n}{p})$ where $n$ is the total number of
786 < sites, and $p$ is the number of processors.  The hull operations
787 < create a set of $p$ hulls each with approximately $\frac{n}{3pr}$
788 < sites (for a cluster of radius $r$).  The communication costs for
789 < distributing this information to all processors is XXX, while the
790 < final computation of the system hull is of order
791 < $O(\frac{n}{3r}\log\frac{n}{3r})$.  Overall, the total costs are
792 < dominated by the computations of the individual hulls, so the Langevin
677 < hull sees roughly linear speed-up with increasing processor counts.
785 > $\mathcal{O}(\frac{n}{p}\log\frac{n}{p})$ where $n$ is the total
786 > number of sites, and $p$ is the number of processors.  These local
787 > hull operations create a set of $p$ hulls, each with approximately
788 > $\frac{n}{3pr}$ sites for a cluster of radius $r$. The worst-case
789 > communication cost for using a ``gather'' operation to distribute this
790 > information to all processors is $\mathcal{O}( \alpha (p-1) + \frac{n
791 >  \beta (p-1)}{3 r p^2})$, while the final computation of the system
792 > hull scales as $\mathcal{O}(\frac{n}{3r}\log\frac{n}{3r})$.
793  
794 + For a large number of atoms on a moderately parallel machine, the
795 + total costs are dominated by the computations of the individual hulls,
796 + and communication of these hulls to create the Langevin Hull sees roughly
797 + linear speed-up with increasing processor counts.
798 +
799   \section*{Acknowledgments}
800   Support for this project was provided by the
801   National Science Foundation under grant CHE-0848243. Computational
802   time was provided by the Center for Research Computing (CRC) at the
803   University of Notre Dame.  
804  
805 + Molecular graphics images were produced using the UCSF Chimera package from
806 + the Resource for Biocomputing, Visualization, and Informatics at the
807 + University of California, San Francisco (supported by NIH P41 RR001081).
808   \newpage
809  
810   \bibliography{langevinHull}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines