cond-mat0601181/man.tex
1: \documentclass[prb,preprint,groupedaddress]{revtex4}
2: %\documentclass[aps,preprint,superscriptaddress]{revtex4}
3: %\documentclass[prb,twocolumn,groupedaddress]{revtex4}
4: % \documentclass[aps,twocolumn,superscriptaddress]{revtex4} 
5: \usepackage{amsmath} 
6: \usepackage{amssymb} 
7: \usepackage{amsfonts}
8: \usepackage[dvips]{graphicx} 
9: %\usepackage[]{caption}
10: \usepackage[]{epsfig} 
11: \bibliographystyle{apsrev} 
12: \voffset 1.5cm
13: \begin{document}
14: %\bibliographystyle{apsrev} 
15: %\bibliographystyle{prsty}
16: 
17: \title{Coupling nonpolar and polar solvation free energies in implicit solvent models}
18:  \author{J. Dzubiella} \email[e-mail address:]
19: {jdzubiella@ucsd.edu} \affiliation{NSF Center for Theoretical
20: Biological Physics (CTBP),} \affiliation{Department of Chemistry and
21: Biochemistry, University of California, San Diego, La Jolla,
22: California 92093-0365} \author{J.~M.~J. Swanson} \affiliation{NSF Center for
23: Theoretical Biological Physics (CTBP),} \affiliation{Department of
24: Chemistry and Biochemistry, University of California, San Diego, La
25: Jolla, California 92093-0365} \author{J.~A. McCammon} \affiliation{NSF
26: Center for Theoretical Biological Physics (CTBP),}
27: \affiliation{Department of Chemistry and Biochemistry, University of
28: California, San Diego, La Jolla, California 92093-0365}
29: 
30: \date{\today}
31: 
32: \begin{abstract}
33: Recent studies on the solvation of atomistic and nanoscale solutes
34: indicate that a strong coupling exists between the hydrophobic,
35: dispersion, and electrostatic contributions to the solvation free
36: energy, a facet not considered in current implicit solvent models. We
37: suggest a theoretical formalism which accounts for coupling by
38: minimizing the Gibbs free energy of the solvent with respect to a
39: solvent volume exclusion function. The resulting differential equation
40: is similar to the Laplace-Young equation for the geometrical
41: description of capillary interfaces, but is extended to microscopic
42: scales by explicitly considering curvature corrections as well as
43: dispersion and electrostatic contributions.  Unlike existing implicit
44: solvent approaches, the solvent accessible surface is an output of our
45: model. The presented formalism is illustrated on spherically or
46: cylindrically symmetrical systems of neutral or charged solutes on
47: different length scales.  The results are in agreement with
48: computer simulations and, most importantly, demonstrate that our
49: method captures the strong sensitivity of solvent expulsion and
50: dewetting to the particular form of the solvent-solute interactions.
51: \end{abstract}
52: %\pacs{61.20.Ja, 68.08.Bc, 87.16.Ac, 87.16.Uv}
53: \maketitle
54: \section{Introduction}
55: Implicit solvent models are widely used in theoretical chemistry to
56: study the solvation of biomolecular systems, as well described in the
57: review of Roux and Simonson\cite{roux:biochem}. They provide a more
58: efficient, although generally less accurate, alternative to
59: atomistically-resolved explicit solvent simulations.  The solvation
60: free energy in these models is usually split into nonpolar (np) and
61: polar (p) terms,
62: \begin{eqnarray}
63: \Delta G=\Delta G_{\rm np}+\Delta G_{\rm p},
64: \end{eqnarray}
65: which are treated in separate energetic evaluations. The nonpolar term
66: includes   the   energetic   cost   of   cavity   formation,   solvent
67: rearrangement,  and solute-solvent  dispersion interactions introduced
68: when  the uncharged  solute is  brought from  vacuum into  the solvent
69: environment. The polar term describes  the free energy of charging the
70: mono- or multipolar solute in the dielectric medium.
71: 
72: The nonpolar term is commonly approximated by surface area models,
73: i.e.~$\Delta G_{\rm np}\simeq\gamma S$, where $S$ is the
74: solvent accessible surface area\cite{richards} and $\gamma$ is an
75: energy per surface area constant, which is a priori not known but
76: fit to atomistic simulations.  The deficiencies of this simple surface
77: area approach have been recognized and a further decomposition of the
78: nonpolar term into cavity (cav) and van der Waals dispersion (vdW) terms has been
79: proposed,\cite{gallicchio:jpcb,gallicchio:jcc} $\Delta G_{\rm
80: np}=\Delta G_{\rm cav}+\Delta G_{\rm vdW}$. This approach has shown
81: improved results for the solvation of
82: alkanes,\cite{gallicchio:jpcb,zacharias} the alanine
83: peptide,\cite{su:biochem} and nonpolar native and misfolded
84: proteins.\cite{levy:jacs} The electrostatic (polar) contribution of
85: the solvation free energy is often approximated by generalized
86: Born\cite{bashford} (GB) or Poisson-Boltzmann\cite{sharp:honig} (PB)
87: models. Both methods use a position-dependent dielectric
88: constant,\cite{roux:biochem} assigned on the basis of
89: the solute surface, which  can be defined in several ways,\cite{baker}
90: or defined implicitly by integration methods.
91: It has been emphasized that all three contributions, $\Delta G_{\rm
92: cav}$, $\Delta G_{\rm vdW}$, and $\Delta G_{\rm p}$ depend critically
93: on the location of the solvent-solute interface.  It has {\it also}
94: been shown that the effective location of the solvent-solute interface
95: can vary according to the local electrostatic\cite{nina}
96: and dispersion\cite{huang:jpcb:2002} potentials. This suggests
97: that interfacial, dispersion and electrostatic contributions should be
98: coupled in implicit solvent approaches. The importance of capturing
99: the right balance between nonpolar and electrostatic contributions
100: in implicit solvation models was emphasized by Ashbaugh
101: {\it et al.} in their study of amphiphiles.\cite{ashbaugh:biophys}
102: 
103: The significance of nonpolar and polar coupling becomes even
104: more evident when solvation is studied on length scales which are
105: large compared to the solvent molecule (typically $\gtrsim$ 1 nm for
106: water), where solvent dewetting ('drying') can occur. In this
107: mechanism, first envisioned by Stillinger,\cite{stilinger} the solvent
108: molecules tend to move away from the surface of a large nonpolar
109: solute forming a liquid-gas like interface parallel to the solute
110: interface. When the surfaces of two
111: large solutes come together dewetting can be amplified due to the
112: gain of interfacial free energy (by decreasing the total
113: liquid-vapor interface area) giving rise to a strong effective
114: attraction.\cite{lum:jpc,hummer:jcp:1997,chandler:review} Early evidence of 
115: confinement-induced dewetting was given only by explicit water
116: simulations for smooth plate-like solutes with a purely repulsive
117: solute-solvent interaction.\cite{wallquist:jpc} More recently,
118: however, it has been demonstrated in varying degrees in several systems with attractive
119: solute-solvent interactions including smooth parallel plate-like
120: solutes,\cite{huang:pnas} atomistically resolved
121: paraffin-plates,\cite{huang:jpcb} graphite-plates\cite{pettitt}, carbon
122: nanotubes,\cite{hummer:nature} and hydrophobic ion channels.\cite{allen:prl,beckstein:jpc,sukharev}
123: 
124: Several of these studies indicated that the magnitude of dewetting is
125: sensitive to the nature of the solute-solvent attractive dispersion
126: interactions.  \cite{huang:pnas,huang:jpcb,pettitt} A similar
127: sensitivity was found in systems where the solutes carry charges or
128: are exposed to an external electric field, e.g. electrostatic interactions have
129: been shown to strongly affect the dewetting behavior of hydrophobic
130: channels\cite{vaitheesvaran, dzubiella:channel1,dzubiella:channel2}
131: and hydrophobic spherical nanosolutes.\cite{dzubiella:jcp:2003,
132: dzubiella:jcp:2004} Furthermore, two recent simulations of proteins
133: supported the importance of solvent dewetting and its sensitivity in
134: realistic biomolecular systems. First, a simulation of the two-domain
135: BphC enzyme showed that the region between the two domains was
136: completely dewetted when vdW and electrostatic interactions were
137: turned off, but accommodated 30$\%$ of the bulk density with the
138: addition of vdW attraction (water was found mainly at the edges of the
139: considered volume, while the central region was still empty), and
140: 85-90$\%$ with the addition of electrostatic
141: interactions.\cite{zhou:science} Second, Liu {\it et al.}  observed a
142: clear dewetting transition in the simulation of the collapse of the
143: melittin tetramer, which was strongly sensitive to the type and
144: location of the hydrophobic residues around the dewetted
145: region.\cite{berne:nature}
146:  
147: Considering the aforementioned studies, we postulate that coupling of
148: the nonpolar and polar solvation contributions in implicit solvent
149: models is crucial for an accurate determination of solvation free
150: energies without too many system-dependent fit parameters. We suggest
151: a general theoretical formalism in which the particular energetic
152: contributions are coupled.  Similar to the approach of Parker {\it et
153: al.}  in their study of bubble formation at hydrophobic
154: surfaces,\cite{attard} we express the Gibbs free energy as a {\it
155: functional} of the solvent volume exclusion function,\cite{beglov:jcp}
156: and obtain the optimal solute surface via minimization.  As
157: we will show, this minimization leads to an expression which is
158: similar to the Laplace-Young equation for the description of
159: macroscopic capillary interfaces,\cite{kralchevsky} but is generalized
160: to explicitly include curvature corrections and solvent-solute
161: interactions, i.e.  short-range repulsion (excluded volume),
162: dispersion, and electrostatics.  This extension of the Laplace-Young
163: theory allows a geometric description of solvation on mesoscopic and
164: microscopic scales. Related approaches in other fields are the
165: Helfrich description of vesicle and membrane
166: surfaces,\cite{helfrich,helfrich2} wetting in colloids and granular
167: media,\cite{kralchevsky,bieker} and functional treatments of
168: electrowetting.\cite{electrowetting}
169: 
170: While most implicit solvent approaches define the solute surface with
171: a geometrical evaluation of the molecular surface, vdW surface, or
172: canonical solvent accessible surface (SAS),\cite{richards,baker} it is
173: an output of our theory.  The surface obtained by minimizing our free
174: energy functional will, in general, be very different than the
175: aforementioned established surface definitions. In particular, our
176: solvent accessible surface should not be confused with the canonical
177: SAS,\cite{richards} which is simply the envelope surrounding
178: probe-inflated spheres. Similarly, phenomenological continuum theories
179: applied to solvent dewetting always assume a certain, simplified
180: geometry for the dry region, e.g. a cylindrical volume for system like
181: hydrophobic ion channels,
182: \cite{allen:jcp,beckstein:jpc,dzubiella:channel2} plate-like
183: particles,\cite{lum:jpc,huang:pnas} or two hydrophobic spherical
184: solutes.\cite{attard2} For a few simple systems this might be a valid
185: approximation but for more complicated solute geometries the shape of
186: the dewetted volume is unknown and a different approach, as suggested
187: in this work, is necessary. We expect our formalism to be particularly
188: useful in solvation studies of large protein assemblies where the
189: hydrophobic surfaces are highly irregular and laced with hydrophilic
190: units,\cite{gerstein,rossky:nature} and for which a unified
191: description of hydration on different length scales is
192: important.\cite{chandler:review} Another potential application is the
193: solvation of superhydrophobic nanosolutes\cite{super} and
194: wetting/dewetting in near-critical colloidal mixtures.\cite{bieker}
195: 
196: A brief summary of our work has been published elsewhere.\cite{prl}
197: Here we present more challenging test cases and an expanded discussion
198: of the approximations and limitations of this model.  The rest of the
199: paper is organized as follows: In section II we present our
200: theoretical formalism and chosen approximations. In section III we
201: first verify that our method can describe solvation on molecular
202: scales with noble gases, ions, and small alkanes. We then demonstrate
203: that it captures the strong sensitivity of dewetting and hydrophobic
204: hydration to specific solute-solvent interactions on larger scales
205: with two alkane-assembled spheres. In section IV we conclude with some
206: final remarks.
207: 
208: \section{Theory}
209: 
210: \subsection{Basic formalism}
211: 
212: Let us consider an assembly of solutes
213: with arbitrary shape and composition surrounded by a dielectric
214: solvent in a volume $\cal W$.  Furthermore, we define a subvolume (or
215: cavity) $\cal V$ empty of solvent for which we can assign a volume
216: exclusion function given by
217: \begin{eqnarray}
218:  v(\vec r)= \begin{cases} 0 & {\text {for}} \,\,\vec r \in \cal V; \cr 1 & {\rm else}. \cr
219:  \end{cases}
220: \end{eqnarray}
221: We assume that the surface surrounding the volume is continuous and
222: closed, i.e.~has no boundary.  The absolute volume $V$ and surface area $S$
223: of $\cal V$ can then be expressed as functionals of $v(\vec r)$ via
224: \begin{eqnarray}
225: V[v]=\int_{\cal W}{\rm d}^3r \;[1-v(\vec r)] \nonumber \\
226: S[v]=\int_{\cal W}{\rm d}^3r \;|\nabla v(\vec r)|,
227: \end{eqnarray}
228: where $\nabla\equiv\nabla_{\vec r}$ is the usual gradient operator
229: with respect to the position vector $\vec r$ and $|\nabla v(\vec r)|$
230: gives a $\delta$-function-like contribution only at the volume
231: boundary. The expression ${\rm d}^3r \;|\nabla v(\vec r)|\equiv{\rm
232: d}S$ can thus be identified as the infinitesimal surface element. In
233: this continuum solvent model the solvent density distribution
234: is simply $\rho(\vec r)=\rho_0 v(\vec r)$, where $\rho_0$ is the
235: bulk density of the solvent at the desired temperature and
236: pressure. Local inhomogeneities of the solvent density, apart from the
237: zero to $\rho_0$ transition at the volume boundaries, are
238: neglected. The solutes' positions and conformations are fixed such that
239: the solutes can be considered as an external potential to the solvent
240: without any degrees of freedom.
241: 
242: As motivated before, we suggest expressing the Gibbs free energy
243: $G[v]$ as a functional of the volume exclusion function
244: $v(\vec r)$, and obtaining the optimal solute volume via
245: minimization
246: \begin{eqnarray}
247: \delta G[v]/\delta v(\vec r)=0,
248: \label{min}
249: \end{eqnarray}
250: where $\delta../\delta v$ denotes the functional derivative with respect to the function $v$. We adopt following
251: ansatz for the Gibbs free energy of the solvent:
252: \begin{eqnarray}
253: \label{eq:grand}
254: &G[v]&=G_{\rm pr}[v]+ G_{\rm int}[v]+ G_{\rm ne}[v]+G_{\rm es}[v]\\
255: &=& P V[v]+\int_{\cal W}{\rm d}^3r\;\gamma(\vec r;[v])|\nabla v(\vec r)|\nonumber \\
256: &+&\rho_0\int_{\cal W}{\rm d}^3r\;v(\vec r) U(\vec r)\nonumber \\ 
257: &+&\!\!\!\!\!\!\int_{\cal W}\!\!\!{\rm d}^3r\!\left\{\frac{\epsilon_0}{2}\epsilon(\vec r;[v])[\nabla\Psi(\vec r)]^2\!-\lambda(\vec r)\Psi(\vec r)\!+v(\vec r)U_{\rm mi}(\vec r)\right\}\nonumber
258: \end{eqnarray}
259: Let us discuss each of the terms in Eq.~(\ref{eq:grand}) in turn.  The
260: first term, $G_{\rm pr}[v]$, proportional to the volume $V$, is the
261: energy of creating a cavity in the solvent against the difference in
262: bulk {\it pressure} between the liquid and vapor phase, $P=P_l-P_v$.
263: For water in ambient conditions, which is close to the liquid-vapor
264: transition, this term is relatively small and can generally be
265: neglected for solutes on molecular scales. The second term $G_{\rm
266: int}[v]$ describes the energetic cost due to solvent rearrangement
267: around the cavity {\it interface} with area $S$ in terms of a free
268: energy/surface area functional $\gamma(\vec r;[v])$. This interfacial
269: energy penalty is thought to be the main driving force behind
270: hydrophobic phenomena.\cite{chandler:review} $\gamma$ is a solvent
271: specific quantity that also depends on the particular topology of the
272: cavity-solvent interface, i.e.~it varies locally in space and is a
273: functional of the volume exclusion function $\gamma=\gamma(\vec
274: r;[v])$.\cite{zwanzig} The exact form of this functional is not known.
275: 
276: The third term, $G_{\rm ne}[v]$, is the total energy of the {\it
277: non-electrostatic} solute-solvent interaction given a solvent density
278: distribution $\rho_0v(\vec r)$. The potential $U(\vec r)=\sum_{i}
279: U_i(\vec r-\vec r_i)$ is the sum of the (short-ranged) repulsive
280: exclusion and (long-ranged) attractive dispersion interaction between
281: each solute atom $i$ at position $\vec r_i$ and a solvent molecule at
282: $\vec r$. Classical solvation studies typically represent the
283: interaction $U_i$ as an isotropic Lennard-Jones (LJ) potential,
284: \begin{eqnarray}
285:  U_{\rm
286: LJ}(r)=4\epsilon\left[\left(\frac{\sigma}{r}\right)^{12}-\left(\frac{\sigma}{r}\right)^6\right],
287: \label{lj}
288: \end{eqnarray} 
289: with an energy scale $\epsilon$, length scale $\sigma$, and
290: center-to-center distance $r$. Using the form of (\ref{lj}) implies
291: that $v({\vec r})$ is defined with respect to the LJ-centers of the
292: solvent molecules.
293: 
294: The last term, $G_{\rm es}[v]$, describes the total energy of the {\it
295: electrostatic} field and the mobile ions in the system expressed by
296: the local electrostatic potential $\Psi(\vec r)$ assuming linear
297: response of the dielectric solvent.\cite{jackson} Similar to
298: $\gamma(\vec r;[v])$, the position-dependent dielectric constant
299: $\epsilon(\vec r)=\epsilon(\vec r;[v])$ depends on the geometry of
300: $v(\vec r)$ with an unknown functional form. $\lambda(\vec r)$ is the
301: fixed charge density distribution of the solutes and the local energy
302: density of the {\it mobile ions} is \cite{sharp:honig,gilson}
303: \begin{eqnarray}
304: U_{\rm mi}(\vec r) = k_BT\sum_{j}\rho_j\{\exp[-\beta q_j \Psi(\vec
305: r)]-1\}.
306: \label{pb}
307: \end{eqnarray}
308: with the thermal energy $k_BT=\beta^{-1}$.  Variation of
309: (\ref{eq:grand}) for a fixed $v(\vec r)$ with respect to $\Psi(\vec
310: r)$ yields the Poisson-Boltzmann equation\cite{sharp:honig,gilson}
311: \begin{eqnarray}
312: {\rm PB}(\vec r)=0 &=& \nabla\cdot[\epsilon_0\epsilon(\vec r;[v])\nabla\Psi(\vec
313: r)]+\lambda(\vec r)\nonumber\\&+&v(\vec r)\sum_{j}q_j\rho_j\exp[-\beta q_j
314: \Psi(\vec r)],
315: \label{pb}
316: \end{eqnarray}
317: where $q_j$ and $\rho_j$ are the charge and concentration of the
318: mobile ion species $j$. Note that the ionic charge density in
319: (\ref{pb}) is multiplied by $v(\vec r)$ to account for the fact that
320: ions usually cannot penetrate the volume empty of polar solvent due to
321: a huge free energy penalty. We remark that the treatment of the
322: electrostatics in our theory has the same limitations as other
323: implicit models using PB, for instance when describing highly charged
324: or strongly correlated electrolyte systems. In contrast to PB/SA
325: models however, the dielectric boundary is optimized such that it
326: responds to the local nonpolar and polar potential; it is not assumed
327: beforehand.
328: 
329: 
330: Let $v_{\rm min}(\vec r)$ be the exclusion function which minimizes
331: the functional (\ref{eq:grand}).  Then, the resulting Gibbs free
332: energy of the system is given by $G[v_{\rm min}]$. The solvation free
333: energy $\Delta G$ is the reversible work to solvate the solute and is
334: given by
335: \begin{eqnarray}
336: \Delta G=G[v_{\rm min}]-G_0,
337: \end{eqnarray}
338: where $G_0$ is a constant reference energy which can refer to the pure
339: solvent state and an unsolvated solute.  The potential of mean force
340: (pmf) along a given reaction coordinate $x$ (e.g. the distance
341: between two solutes' centers of mass) is given, within a constant, by
342: $G[v_{\rm min}]$, where $v_{\rm min}(\vec r)$ must be evaluated for
343: every $x$.  In order to proceed we will need valid approximations for
344: $\gamma(\vec r;[v])$ and $\epsilon(\vec r;[v])$ with which $v_{\rm
345: min}(\vec r)$ can be calculated by explicitly minimizing our free
346: energy functional (\ref{eq:grand}) according to (\ref{min}).
347: 
348: \subsection{Approximations for $\gamma(\vec r;[v])$ and $\epsilon(\vec r;[v])$}
349: 
350: Let us start with a possible description of $\gamma(\vec r;[v])$.  For
351: a planar macroscopic interface the parameter $\gamma$ is usually
352: identified by the surface tension of the solvent adjacent to the
353: second medium. This surface tension obviously depends on the
354: microscopic interactions between the medium and the solvent and is
355: generally decreased by attractive dispersion or electrostatic
356: contributions.  It seems that microscopic interactions are adequately
357: represented by a macroscopic quantity like $\gamma$ {\it if} their
358: range is much smaller than the investigated length scales, such as the
359: radii of curvature or mean particle distances. The effect of the
360: microscopic interactions are then absorbed in $\gamma$. This has been
361: exemplified with free energy estimates for the solvation of large,
362: neutral plate-like or spherical alkane-assembled
363: solutes,\cite{huang:pnas,huang:jpcb:2002} For the description of
364: solvation on smaller length scales, however, it seems important to
365: separate the free energy into a part which accounts for the formation
366: of a cavity and a part which describes the dispersion interactions
367: explicitly.\cite{gallicchio:jpcb} Furthermore, it has been shown that
368: the water liquid-vapor surface tension $\gamma_{\rm lv}$ is the
369: asymptotic value of the solvation free energy per surface area for
370: hard spherical cavities in water in the limit of large
371: radii.\cite{lum:jpc,huang:jpc} These considerations motivate our
372: choice of the second and third term in the functional (\ref{eq:grand})
373: and lead to the assumption $\gamma\equiv\gamma_{\rm lv}$ in the limit
374: of vanishing curvatures.
375: 
376: The surfaces of realistic (bio)molecules, however, display highly
377: curved shapes, so $\gamma(\vec r;[v])$ will strongly depend on the interface
378: geometry around $\vec r$ in a complicated fashion. In the following we
379: make a {\it local curvature approximation}, i.e.~we assume that
380: $\gamma(\vec r;[v])$ can be expressed solely as a function of the local mean
381: curvature
382: \begin{eqnarray}
383: H(\vec r)=\left(\kappa_1(\vec r)+\kappa_2(\vec r)\right)/2=R(\vec
384: r)^{-1},
385: \end{eqnarray}
386: where $R(\vec r)$ is the radius of mean curvature and $\kappa_1(\vec r)$ and
387: $\kappa_2(\vec r)$ are the local principal curvatures of the
388: interface.\cite{prince} The mean curvature $H$ is only defined at the
389: boundary of $v(\vec r)$. We have chosen the convention in which the
390: curvatures are positive for convex surfaces (e.g. a spherical cavity)
391: and negative for concave surfaces (e.g. a spherical droplet).
392: 
393: The curvature dependence of the liquid-vapor surface tension
394: has been a long standing subject of research and is still under steady
395: discussion.\cite{hendersonreview,bieker,evans2} For water, which is
396: close to the critical point under ambient conditions, $\gamma$ is
397: argued to be nonanalytical in curvature.\cite{evans2}
398: The first order correction term, however, is likely to be linear in
399: curvature as predicted by scaled-particle theory,\cite{spt} the
400: commonly used ansatz to study the solvation of hard spherical
401: cavities. Although this result is only strictly valid for the case of
402: spherical particles, we assume that it can be applied to local mean
403: curvatures such that $\gamma(\vec r;[v])$ reduces to the function
404: \begin{eqnarray}
405: \gamma(\vec r)=\gamma_{{\rm lv}}(1-2\delta H(\vec r)),
406: \label{H}
407: \end{eqnarray}
408: where $\delta$ is the Tolman length, which is expected to be of
409: molecular size.\cite{tolman} In our study we assume $\delta$ is
410: constant and positive, while the curvature can be positive or negative
411: as defined above.  Note that this leads to an increase of surface
412: tension for concave surfaces in agreement with the geometrical
413: arguments of Honig {\it et al.}\cite{nicholls} in their solvation
414: study of alkanes. It has been shown by computer simulations of growing
415: a hard spherical cavity in water that (\ref{H}) predicts the
416: interfacial energy rather well for radii $\gtrsim
417: 3$\AA.\cite{huang:jpc} A major drawback of approximation (\ref{H}) is
418: that it gives unphysical results if the radius of mean curvature is
419: smaller than twice the Tolman length, $|R|<2\delta$. It yields
420: negative and diverging surface tensions for convex and concave
421: surfaces, respectively.  The latter is not possible due to the finite size
422: of the solvent molecules. Thus, care has to be taken with approximation (\ref{H}) when investigating systems which can exhibit radii of
423: curvature $|R|<2\delta$.
424: %% We note that for small convex
425: %% cavities it has been shown that Gaussian fluctuations of the solvent
426: %% dominate the solvation free energy and the latter can be expressed
427: %% analytically in terms of the cavity
428: %% volume.\cite{hummer:pnas,garde:prl} An improved approximation for
429: %% $\gamma(v)$ (or in general for $G[v]$) incorporating this behavior may
430: %% be possible and needs further investigation.
431: %% Another option or additional
432: %% refinement may be the inclusion of a second order correction
433: %% proportional to the curvature squared as suggested by Helfrich, such
434: %% that
435: %% \begin{eqnarray}
436: %% \gamma(H)=\gamma_{{\rm lv}}(1+2\delta H+aH^2+bK),
437: %% \label{helf}
438: %% \end{eqnarray}
439: %% where $a$ and $b$ are a priori unknown rigidity constants, and
440: %% $K=\kappa_1\kappa_2$ is the Gaussian
441: %% curvature.\cite{helfrich,helfrich2} Using (\ref{helf}) in the
442: %% functional (\ref{eq:grand}) without excluded volume, dispersion and
443: %% electrostatic terms leads to the Helfrich
444: %% functional,\cite{helfrich,helfrich2} almost exclusively used for the
445: %% shape description of membranes and vesicles.
446: 
447: Let us now turn to electrostatics. The most common approximation for the
448: position-dependent dielectric constant is proportional to the volume
449: exclusion function $v(\vec r)$,\cite{roux:biochem} such that the functional
450: $\epsilon(\vec r;[v])$ reduces to,
451: \begin{eqnarray}
452: \epsilon(\vec r)=\epsilon_v+v(\vec r)(\epsilon_l-\epsilon_v),
453: \label{e}
454: \end{eqnarray}
455: where $\epsilon_v$ and $\epsilon_l$ are the dielectric constants
456: inside and outside the volume $\cal V$, respectively.  Eq.~(\ref{e})
457: is valid only in the limit of large solute sizes when the molecular size
458: of the solvent is negligible. For charged solutes on a molecular
459: scale, let's say mono- or polyvalent ions, two difficulties
460: arise. First the electric field close to the highly curved solutes can
461: be strong enough for the dielectric constant to be field
462: dependent. This formally affects the form of the electrostatic term in
463: the free energy functional which assumes a linear response of the
464: solvent. An improvement for continuum models along these lines has
465: been proposed by Luo et al.\cite{luo:jpc} Second, the effective
466: position of the dielectric boundary is known to depend on the sign of
467: the solute charge for asymmetric solvent molecules like water. This
468: expresses itself, for instance, in different Born radii for two
469: equally charged ions which have exact the same LJ parameters but a
470: different sign of charge. A reasonable improvement of (\ref{e}) would
471: be to shift the dielectric boundary at $\vec r$ {\it parallel} to the
472: volume boundary by a potential dependent amount $\Delta \vec r=\xi(\Psi(\vec r))\vec n(\vec r)$:
473: \begin{eqnarray}
474: \epsilon(\vec r)=\epsilon_v+v\left(\vec r-\Delta \vec r\right)(\epsilon_l-\epsilon_v),
475: \label{e2}
476: \end{eqnarray}
477: where $\vec n$ is the unit normal vector to the interface. We do not
478: attempt however, to find an approximation for the function $\xi(\Psi)$
479: in this work and postpone this investigation to later studies. For
480: further illustration of our approach we content ourselves with the
481: approximations (\ref{H}) for $\gamma(\vec r;[v])$ and (\ref{e}) for
482: $\epsilon(\vec r;[v])$.
483: 
484: \subsection{Minimization of the free energy functional}
485: For the functional derivative of the interfacial term,
486: $G_{\rm int}[v]$, we utilize
487: \begin{eqnarray}
488: \frac{\delta}{\delta v}\int_{\cal W}{\rm d}^3r \;|\nabla v(\vec r)|=\frac{\delta}{\delta v}\int_{\partial\cal W}{\rm d}S=-2H
489: \end{eqnarray}
490: and
491: \begin{eqnarray}
492: \frac{\delta}{\delta v}\int_{\cal W}{\rm d}^3r\;H \;|\nabla v(\vec r)|=\frac{\delta}{\delta v}\int_{\partial\cal W}{\rm d}S\;H=-K,
493: \end{eqnarray}
494: which has been derived in detail by Zhong-can and Helfrich by means of
495: differential geometry.\cite{helfrich} The variable
496: $K(\vec r)=\kappa_1(\vec r)\kappa_2(\vec r)$ is the Gaussian curvature of the interface, which
497: is an intrinsic geometric property of $v$. Plugging in approximations
498: (\ref{H}) and (\ref{e}) into (\ref{eq:grand}), and minimizing with
499: (\ref{min}), we obtain,
500: \begin{eqnarray}
501: 0&=&{\rm de}(\vec r)= P+2\gamma_{{\rm lv}}\left[H(\vec r)-\delta K(\vec
502: r)\right]-\rho_0 U(\vec r)\nonumber\\ &-&\frac{\epsilon_0}{2}[\nabla\Psi(\vec
503: r)\epsilon(\vec
504: r)]^2\left(\frac{1}{\epsilon_l}-\frac{1}{\epsilon_v}\right)-U_{\rm
505: mi}(\vec r).\nonumber\\
506: \label{diff}
507: \end{eqnarray}
508: Eq.~(\ref{diff}) is a partial second order differential equation (de)
509: for the optimal exclusion function $v_{\rm min}(\vec r)$ expressed in
510: terms of pressure, curvatures, short-range repulsion, dispersion, and
511: electrostatic terms, all of which have dimensions of energy density.
512: It can also be interpreted as a mechanical balance between the forces
513: per surface area generated by each of the particular contributions.
514: Thus, in our approach the surface shape and geometry, expressed by $H$
515: and $K$, are directly related to the inhomogeneous potential
516: contributions.  The constant solute charge density
517: $\lambda(\vec r)$ does not appear explicitly in (\ref{diff}) but is
518: implicitly considered in the PB equation (\ref{pb}), which must be
519: solved simultaneously. If curvature correction ($K$-term) and the last
520: three energetic terms are neglected one obtains the Laplace-Young
521: equation,
522: \begin{eqnarray}
523: P=-2\gamma_{lv}H,
524: \label{ly}
525: \end{eqnarray}
526: which is exclusively used for the shape description of macroscopic
527: capillary and interfacial phenomena in conjunction with appropriate
528: boundary conditions, e.g. prescribed liquid-solid contact angles at
529: the solid surfaces.\cite{kralchevsky} In our description the boundary
530: conditions are provided by the constraints given by the short-ranged
531: repulsive term in $U(\vec r)$, and the distribution of dispersion and
532: electrostatics, allowing an extrapolation of the Laplace-Young
533: description to mesoscopic and microscopic scales. Notice that in our
534: approach the solvent is treated as a continuum while the solute is
535: explicitly resolved.  One could use a coarse-grained treatment for the
536: solute by including the appropriate non-electrostatic and
537: electrostatic interactions in (\ref{eq:grand}).
538: 
539: The solution of (\ref{diff}) requires an appropriate
540: parametrization, i.e.~coordinate representation, for
541: the curvatures $H$ and $K$, such that the equation is expressed as a
542: function of the vector $\vec r$ and its first and second derivatives
543: in space. Analytical solutions to the much simpler Eq.~(\ref{ly}) and thus
544: to (\ref{diff}) are only available for systems with very simple geometries.\cite{kralchevsky}
545: %% Examples follow in the next section.  It seems that analytical
546: %% solutions to (\ref{diff}) are only available for very simple systems,
547: %% e.g., one or two hard and neutral spheres,\cite{kralchevsky} but
548: %% further investigation is needed. The stability of (\ref{eq:grand})
549: %% should also be checked by verifying that the second variation is
550: %% positive definite,\cite{helfrich2} $\delta^{(2)}G/\delta v^2>0$, but
551: %% seems analytically not easily tractable too.
552: Thus we use numerical solutions of (\ref{diff}) in the following to
553: further illustrate our theory.
554: 
555: \section{Applications}
556: 
557: First, we will consider the solvation of microscopic solutes such as
558: noble gases, simple alkanes, and ions which can be treated as neutral
559: or charged Lennard-Jones spheres. Then, we will investigate alkane
560: assemblies on a larger scale where interfacial and dewetting effects
561: are much more dominant.  For simplicity and a better transparency of
562: the results, mobile ions will be neglected in these illustrations.
563: 
564: \subsection{One Lennard-Jones sphere}
565: 
566: In this section we compare our approach to results from SPC and SPC/E
567: explicit solvent simulations.\cite{berendsen:jpc} We refrain from 
568: comparing to real experiments since approximations in computer
569: experiments are more easily controlled and the LJ parameters
570: of the solutes are commonly parametrized to yield accurate results in
571: classical simulations.
572: 
573: For a spherical solute with a charge $Q$ homogeneously distributed
574: over its surface, the functional (\ref{eq:grand}) with approximations
575: (\ref{H}) and (\ref{e}) and no mobile ions reduces to a function of
576: $R$, the radius of the sphere empty of solvent. The solvation free
577: energy is
578: \begin{eqnarray}
579: \Delta G(R)&=&\Delta G_{\rm pr}(R)+\Delta G_{\rm int}(R)+\Delta G_{\rm ne}(R)+\Delta G_{\rm es}(R)\nonumber\\
580: &=&\frac{4}{3}\pi R^3 P + 4\pi R^2 \gamma_{{\rm
581:  lv}}\left(1-\frac{2\delta}{R}\right)\nonumber\\&+& \int_R^{\infty} 4\pi r^2{\rm
582:  d}r\; \rho_0U_{\rm LJ}(r)\nonumber\\ &+&\frac{Q^2}{8\pi \epsilon_0
583:  R}\left(\frac{1}{\epsilon_l}-\frac{1}{\epsilon_v}\right)
584: \label{eq:sphere}. 
585: \end{eqnarray}
586: Note that the last term in (\ref{eq:sphere}) is equivalent to the Born
587: electrostatic solvation free energy.\cite{roux:biochem} Recently,
588: Manjari {\it et al.} have presented a very similar expression for the
589: solvation of a charged spherical cavity on the basis of a
590: minimization principle and have investigated the variation of $R$ with
591: thermodynamic conditions.\cite{kim} Differentiation of 
592: (\ref{eq:sphere}) with respect to $R$ and subsequent division by $4\pi
593: R^2$ yields
594:  \begin{eqnarray}
595:   0&=& P+\frac{2\gamma_{{\rm
596:   lv}}}{R}\left(1-\frac{\delta}{R}\right)-\rho_0U_{\rm
597:   LJ}(R)\nonumber\\&-&\frac{Q^2}{32\pi^2\epsilon_0
598:   R^4}\left(\frac{1}{\epsilon_l}-\frac{1}{\epsilon_v}\right)
599: \label{sphere2},
600:  \end{eqnarray}
601: which is in accord with Eq.~(\ref{diff}) given sphere-like curvatures,
602: $H=1/R$ and $K=1/R^2$.   We
603: can now calculate the solvation free energies of simple spherical
604: solutes, such as noble gases or ions.  The free parameters in
605: Eq.~(\ref{sphere2}) are the pressure $P$, Tolman length $\delta$,
606: liquid-vapor surface tension $\gamma_{\rm lv}$, and dielectric
607: constants $\epsilon_v$ and $\epsilon_l$.
608:  
609: \subsubsection{One neutral LJ sphere}
610: First, let us focus on uncharged spheres, for which the electrostatic
611: term in (\ref{eq:sphere}) can be neglected.  We compare the results
612: from our theory to those calculated by Hummer {\it et
613: al.}\cite{hummer:jpc:1996} for neutral LJ spheres in SPC water, and
614: those calculated by Paschek\cite{paschek} for noble gases in SPC and
615: SPC/E water.  The solute-water LJ parameters $\sigma$ and $\epsilon$
616: are summarized in Tab.~I.  The surface tension $\gamma_{\rm lv}$ was
617: set to that estimated for SPC and SPC/E water at 300K, $\gamma_{\rm
618: lv}=65$mJ/m$^2$ and $\gamma_{\rm lv}=72$mJ/m$^2$,
619: respectively.\cite{alejandre,huang:jpc} The pressure is fixed to 1atm.
620: Finally, the remaining free parameter $\delta$ was fit to {\it reproduce}
621: the simulation solvation free energies exactly.  The solvation
622: free energies from simulation $\Delta G_{\rm sim}$ and best fit Tolman
623: lengths $\delta_{\rm bf}$ are shown in Tables~I~and~II for the SPC and
624: SPC/E models, respectively.
625: 
626: Before we discuss the results, let us compare the particular energy
627: contributions $\Delta G_{\rm i}(R)$ with ${\rm i=pr,int,ne}$ for
628: Na$^0$ (plotted in Fig.~\ref{fig:G}). As anticipated, the pressure
629: term $\Delta G_{\rm pr}(R)$ with $P=1$atm is negligible compared to
630: the other contributions. The interfacial term $\Delta G_{\rm int}(R)$
631: increases with the cavity radius $R$. The integrated LJ-interaction
632: term $\Delta G_{\rm ne}(R)$ shows long-range attraction and a steep
633: short-ranged repulsion with a minimum at
634: $R=\sigma(Na^{0})=2.85$\AA. The total solvation free energy for the
635: Na$^0$ shows a single minimum at $R_{\rm min}=2.32$\AA~ with $\Delta
636: G=9.2$kJ/mol for a $\delta_{\rm bf}$=0.79\AA.
637: 
638: 
639: The results for the other LJ-spheres, summarized in Tab.~I and II,
640: reveal several noteworthy observations.  First, the best fit Tolman
641: lengths $\delta_{\rm bf}$ range from 0.76\AA~ to 1.00\AA~; they are
642: not only of molecular size, as expected, but are approximately half
643: the LJ-radius of a SPC or SPC/E water molecule.  Second, the
644: $\delta_{\rm bf}$ values for noble gases in SPC/E water (Tab.~II) are
645: approximately 10$\%$ larger than those in SPC water (Tab.~I).  This is
646: in qualitative agreement with the results of Huang {\it et al.} who
647: measured $\delta$=$0.76\pm0.05$\AA~ and $\delta$=$0.90\pm0.03$\AA~ for
648: SPC and SPC/E, respectively, by fitting Eq.~(\ref{H}) to the hydration
649: free energy of hard spheres with varying radii.\cite{huang:jpc}
650: 
651: Third, the quite accurate data of Paschek demonstrate a systematic increase of 
652: $\delta_{\rm bf}$ with solute size.  The inability of our theory 
653: to be fit by one fixed constant $\delta_{\rm bf}$ points to the 
654: anticipated fact that Eq.~(\ref{H}) can not capture strong curvature 
655: effects accurately and will have to be refined for small solutes. Despite this shortcoming,
656: these results show surprisingly good agreement; if we assume a fixed
657: delta, for instance $\delta=0.91$\AA~ for all noble gases in the SPC
658: data of Paschek, our theory predicts results within
659: $15\%$ of the simulation data. Finally, we
660: observe that the effective optimal sphere radius $R_{\rm min}$ is
661: always smaller than the radius of the canonical SAS with a
662: typical probe radius of 1.4\AA,\cite{connolly} $R_{\rm
663: min}<(\sigma_{ss}/2+1.4{\rm \AA})\simeq\sigma$, but larger than the
664: vdW surface, $R_{\rm min}>\sigma_{ss}/2$, where
665: $\sigma_{ss}$ is the solute-solute LJ-length.\cite{sigmanote}
666: 
667: 
668: \subsubsection{One charged LJ sphere}
669: Let us now turn to charged Lennard-Jones spheres (ions) also examined
670: in the paper by Hummer {\it et al.}  with SPC water simulations. We
671: assume $\delta$ to be the {\it fixed} by the previously obtained
672: $\delta_{\rm bf}$ values for uncharged spheres.  The dielectric
673: constants are set to $\epsilon_v=1$ and $\epsilon_l=65$, in accord
674: with SPC water.\cite{spoel} The electrostatic contribution $\Delta
675: G_{\rm es}(R)$ and the total $\Delta G(R)$ for Na$^+$ are shown in the
676: inset of Fig.~\ref{fig:G}. The electrostatic contribution decreases
677: the optimal radius to $R_{\rm min}=1.83$\AA~ giving a solvation free
678: energy of $\Delta G=-334$kJ/mol. In fact, the optimal sphere radius
679: $R_{\rm min}$ is always considerably smaller for the charged solutes
680: (Tab.~III) than for their neutral counterparts (Tab.~I). This is
681: caused by the strong compressing force of the polar solvent attempting
682: to penetrate the low dielectric cavity.  The solvation free energies
683: from theory $\Delta G$ and those from simulation $\Delta G_{\rm sim}$
684: are also shown in Tab.~III.  While our theory describes the hydration
685: free energies for positively charged ions within $15\%$, it
686: considerably underestimates those of the negative ions. This
687: qualitative disagreement between positive and negative ions was
688: expected since the Born radii for anions are always smaller than those
689: for cations, a consequence of the different solvation structure around
690: charged solutes with opposite signs. As mentioned in the previous
691: section, the position of the dielectric boundary has to be refined for
692: accurate estimates of the electrostatic contribution to the hydration
693: free energy. If we apply the correction (\ref{e2}) to the dielectric
694: boundary with a simple, potential-independent shift $\xi_+=-0.25$\AA~
695: for positive and $\xi_-=-1.05$\AA~ for negative spheres such that the
696: dielectric boundary has a radius of $R+\xi_\pm<R$, improved values
697: ($\Delta G_\xi$ in Tab.~III) are obtained which reproduce all
698: simulation values within $10\%$!
699:  
700: 
701: 
702: \subsection{Linear alkanes}
703: 
704: Let us now consider simple polyatomic molecules, such as ethane,
705: propane, or butane in a cylindrically symmetric one-dimensional (1D)
706: chain conformation.  Other conformations will be neglected. The
707: symmetry of these systems allows us to express the volume exclusion
708: function $v(\vec r)$ of the enveloping surface by a one dimensional
709: shape function $r(z)$, where $z$ is the coordinate on the symmetry
710: axis and $r$ the radial distance to it. The full surface in
711: three-dimensional space is obtained by revolving the shape function
712: $r(z)$ around the symmetry axis. Technical details are given in the
713: Appendix.
714: 
715: The LJ parameters for ethane and methane are the same as those used by
716: Ashbaugh {\it et al.}\cite{ashbaugh:biophys} in their SPC simulation
717: of linear alkanes (see Tab.~I).  The simulation solvation energy of
718: the spherical methane, $\Delta G=10.96$kJ/mol, can be reproduced with
719: $\delta_{\rm bf}=0.85$\AA. Solving the cylindrically symmetric problem
720: for ethane using the same $\delta$, we obtain a fit-parameter free
721: $\Delta G=11.40$kJ/mol, which is only 7$\%$ larger than the simulation
722: results. Alternatively, $\delta_{\rm bf}=0.87$\AA~ reproduces the
723: simulation energy exactly.  This is surprisingly good agreement
724: considering the crudeness of our curvature correction and the fact
725: that the large curvature of the system varies locally in space. The
726: curvature and shape functions are plotted in Fig.~\ref{fig:ethane}
727: together with the vdW surface and the canonical SAS obtained from
728: rolling a probe sphere with the typically chosen radius $r_{\rm
729: p}=1.4$\AA~ over the vdW surface.\cite{connolly} Away from the
730: center of mass $|z|\gtrsim 1$\AA~ the curvatures follow the expected
731: trends for the spherical surfaces: $H= 1/R$ and
732: $K=1/R^2$ with $R \simeq 3.1$\AA~.  The
733: optimal surface resulting from our theory is smaller than the
734: canonical SAS and smooth at the center of mass ($z=0$) where the
735: canonical SAS has a kink. Thus our surface has a smaller mean
736: curvature at $z=0$ and an almost zero Gaussian curvature, which is
737: typical for a cylinder geometry with one principal curvature equal to
738: zero.  These results may justify the use of smooth surfaces in
739: coarse-grained models of closely-packed hydrocarbons, a possibility we
740: will explore in the following section with solvation on larger length
741: scales where dewetting effects can occur.  If we repeat the above
742: calculation for propane and butane (three and four LJ-spheres, see
743: also Tab.~I for parameters) we need $\delta_{\rm bf}=0.94$\AA~ and
744: $\delta_{\rm bf}=0.96$\AA, respectively, to reproduce the simulation
745: results exactly. The increasing difference in $\delta_{\rm bf}$
746: compared to methane and ethane is likely due to contributions from
747: other than cylindrically symmetric conformations which were ignored in
748: our analysis.
749:  
750: 
751: \subsection{Two spherical nanosolutes}
752: 
753: \subsubsection{Model}
754: 
755: Let us now consider two spherical solutes which represent
756: homogeneously assembled CH$_2$ groups with a uniform density
757: $\rho$=0.024\AA$^{-3}$ up to a radius $R_0=15$\AA, defined by the
758: maximal distance between a CH$_2$ center and the center of the solute.
759: Integration of the CH$_2$-water LJ interaction over the entire sphere
760: yields a 9-3 like potential $U_i(r)$ for the interaction between the
761: center of the solute ($i=1,2$) and a water
762: molecule. \cite{huang:jpcb:2002} The intrinsic, nonelectrostatic
763: solute-solute interaction $U_{\rm ss}$ can be obtained in a similar
764: fashion.  The CH$_2$-water LJ parameters, $\epsilon=0.5665$kJ/mol and
765: $\sigma=3.52$\AA, are taken from the OPLSUA
766: force-field\cite{jorgensen} and are similar to those used by Huang
767: {\it et al.} in their study on dewetting between paraffin
768: plates.\cite{huang:jpcb} Minimizing Eq.~(\ref{eq:sphere}) for just one
769: sphere we obtain an optimal solvent excluded radius of $R_{\rm min}
770: \simeq 17.4$\AA, which is $R_0+2.4$\AA. Since we are also
771: interested in the effects of electrostatic interactions we place
772: opposite charges $\pm Ze$, where $e$ is the elementary charge, in the
773: center or on the edge of the two spheres. Poisson's equation is
774: simultaneously solved on a two-dimensional grid in cylindrical
775: coordinates. Numerical details are given in the Appendix.
776: 
777: The solvation of the two solutes is studied for a fixed
778: surface-to-surface distance which we define as $s_0=r_{12}-2R_0$,
779: where $r_{12}$ is the solute center-to-center distance. The effective
780: surface-to-surface distance defined by the accessibility of the
781: solvent centers is thus $s\simeq r_{\rm 12}-2R_{\rm
782: min}=s_0-4.8$\AA. In the following we focus on a separation distance
783: of $s_0=8$\AA~ to investigate the influence of different energetic
784: contributions on the shape function, $r(z)$, and the curvatures,
785: $K(z)$ and $H(z)$. For $s_0 = 8$\AA, it follows that $s\simeq 3.2
786: $\AA, such that two water molecules could fit between the solutes on
787: the $z$-axis. We systematically change the solute-solute and
788: solute-solvent interactions, as summarized in Tab.~I. We begin with
789: only the repulsive part of the nonelectrostatic interaction $U_i(r)$
790: in system I, and then add a curvature correction with
791: $\delta=0.75$\AA, vdW attractions, and sphere-centered charges $Z = 4$
792: and $Z = 5$ in systems II-V, respectively.  To study the influence of
793: charge location, we reduce the magnitude of each charge in system VI
794: to $Z=1$ and move them to the edge of the spheres on the symmetry axis
795: such that they are $8$\AA~apart (indicated by arrows in
796: Fig.~\ref{fig:nano}.  The surface tension and dielectric constants of
797: the vapor (solute) and liquid are fixed to $\gamma_{\rm
798: lv}=72$mJ/m$^2$, $\epsilon_v=1$, and $\epsilon_l=78$, respectively.
799: 
800: \subsubsection{Behavior of the shape function}
801: 
802: 
803: 
804: 
805: The results for the curvatures and shape function, defined by $r(z)$,
806: for systems I-VI are shown in Fig.~\ref{fig:nano}. Away from the
807: center of mass ($|z|\gtrsim 10$\AA), systems I-VI show very little
808: difference. The curvatures are $H=1/R$ and $K=1/R^2$ with $R \simeq
809: 17.4$\AA. Close to the center of mass ($z\simeq 0$), however, the
810: influence of changing the parameters is considerable. In system I,
811: Eq.~(\ref{diff}) reduces to the minimum surface equation $H(z)=0$ for
812: $z\simeq 0$. For two adjacent spheres the solution of this equation is
813: the catenoid $r(z)\simeq{\rm cosh}(z)$, which features zero mean
814: curvature ($\kappa_1$ and $\kappa_2$ cancel each other) and negative
815: Gaussian curvature. As a consequence, the system exhibits a vapor
816: bubble bridging the solutes, i.e. water is removed from the region
817: between the spheres even though it fits there. This dewetting is
818: driven by the interfacial term $G_{\rm int}$ which always favors
819: minimizing the liquid-vapor interface.
820: 
821: When curvature correction is applied (system II), the mean curvature
822: becomes nonzero and negative (concave) at $z\simeq 0$, while the
823: Gaussian curvature grows slightly more negative. Thus the total
824: enveloping surface area becomes larger and the solvent inaccessible
825: volume shrinks, i.e. the value of the shape function at $z \simeq 0$
826: decreases. Turning on solute-solvent dispersion attraction amplifies
827: this trend significantly as demonstrated by system III. Mean and
828: Gaussian curvatures increase fivefold, showing strongly enhanced
829: concavity, and the volume empty of water decreases considerably,
830: expressed by $r(z=0)\simeq 10.7$\AA~ dropping to $r(z=0)\simeq
831: 6.3$\AA. These trends continue with the addition of electrostatics in
832: system IV.  When the sphere charges are further increased from $Z=4$
833: to $Z=5$ (system IV$\rightarrow$V), we observe a wetting transition:
834: the bubble ruptures and the shape function jumps to the solution for
835: two isolated solutes, where $r(z\simeq 0)=0$. The same holds when
836: going from III to VI, when only one charge unit, $Z=1$, is placed at
837: each of the solutes' surfaces. Importantly, this demonstrates that the
838: present formalism captures the sensitivity of dewetting phenomena to
839: specific solvent-solute interactions as reported in previous
840: studies.\cite{huang:jpcb,pettitt,dzubiella:channel1,vaitheesvaran,dzubiella:jcp:2003,zhou:science,berne:nature}
841: Note that the optimal shape function at $|z|\simeq\pm2$\AA~ is closer
842: to the solutes in VI compared to V due to the proximity of the charge
843: to the interface. Clearly, the observed effects, particularly the
844: transition from III to VI, cannot be described by existing solvation
845: models which use the surface area (GB/SA or PB/SA)\cite{roux:biochem}
846: or effective surface tensions and macroscopic solvent-solute contact
847: angles\cite{kralchevsky,attard} as input.
848: 
849: \subsubsection{Potential of mean force}
850: The significant change of the shape function with the
851: solute-solvent interaction has a strong impact on the potential of
852: mean force (pmf) (or effective interaction) between the solutes
853: \begin{eqnarray}
854: W(s_0)= G(s_0)- G(\infty)+U_{\rm ss}(s_0).
855: \end{eqnarray} 
856: Recall that $U_{\rm ss}$ is the instrinsic dispersion interaction
857: between the two solutes. Values of $W(s_0=8{\rm\AA})$ are given in
858: Tab.~IV. From system I to VI the total attraction between the solutes
859: decreases almost two orders of magnitude. Interestingly, the curvature
860: correction (I$\rightarrow$II) lowers $W$ by a large 23.5$k_BT$, even
861: though $R_0\gg\delta$. The reason is that the mean radii of curvature
862: between the spheres can assume values $\simeq \delta$, implying that
863: curvature correction is also important for large solutes. A striking
864: effect occurs when vdW contributions are introduced
865: (II$\rightarrow$III): the inter solute attraction decreases by
866: approximately $28k_BT$ while the dispersion solute-solute potential
867: $U_{\rm ss}(s_0=8{\rm \AA})$ changes by only $-0.44 k_BT$. Similarly,
868: adding charges of $Z=5$ (III $\rightarrow$ V) at the solutes' centers
869: or $Z=1$ (III $\rightarrow$ VI) at the solutes' surfaces decreases the
870: total attraction by 1.2$k_BT$ and 5k$_B$T, respectively. Note that the
871: total attraction {\it decreases} even though electrostatic attraction
872: has been added between the solutes. The same trend has been observed
873: recently in explicit water simulations of a similar system of charged
874: hydrophobic nanosolutes.\cite{dzubiella:jcp:2003,dzubiella:jcp:2004}
875: 
876: Now we turn our attention to varying the intersolute distance.  The
877: pmfs and solute-solute mean forces $F=\partial W(s_0)/\partial s_0$
878: versus a range of $s_0$ are shown for systems I,II,III, and VI in 
879: Fig.~\ref{fig:pmf}. System I, with purely repulsive
880: solute-solvent interactions, displays a strong attraction
881: ($W\simeq-150k_BT$) at $s_0=2$\AA~ which decreases, almost linearly,
882: to zero at a distance $s_0=13.5$\AA~ where the system shows a wetting
883: transition. The corresponding force is discontinuous at this critical
884: distance.  The steep repulsion at short intersolute distances
885: ($s_0\simeq 1.5$\AA~) stems from the repulsive term of the LJ
886: interaction between the solutes. Note that the intrinsic solute-solute
887: interaction $U_{\rm ss}(s_0)$, also shown in Fig.~\ref{fig:pmf}, is almost two
888: orders of magnitude smaller than the hydrophobic attraction. Adding
889: the curvature correction in system II decreases the range and strength
890: of the pmf by approximately $20\%$, which is significant and
891: unexpected since $R_0\gg\delta$.  Adding dispersion
892: attractions in system III decreases the range and strength of the
893: hydrophobic attraction considerably, but it is still much stronger
894: than the inter solute dispersion attraction $U_{ss}$ alone. When
895: surface charges ($Z=1$) are added in system VI, the range of
896: hydrophobic attraction further decreases but the total attraction
897: increases at short intersolute distances.  This is due to the
898: increasing size of the bridging bubble ($r(z=0)$ increases) as the two
899: solutes approach each other, which decreases the high dielectric
900: screening of the solute-solute electrostatic attraction.  This again
901: underlines the importance of coupling electrostatics and dewetting
902: effects, as the electrostatic attraction (or repulsion) may be
903: magnified by more than an order of magnitude when dewetting
904: occurs. For charges with opposite sign this could be interpreted as the
905: stabilization of a salt bridge due to dehydration.\cite{fernandez}
906: Systems IV and V, not shown in Fig.~\ref{fig:pmf}, exhibit the same
907: qualitative behavior as system VI.
908: 
909: 
910: \subsubsection{Comparison of mean forces to previous MD simulations}
911: We continue considering the mean force between two nanosized solutes
912: and compare our theory now to the MD simulations of Dzubiella {\it et
913: al.}\cite{dzubiella:jcp:2003,dzubiella:jcp:2004} Their solute model
914: slightly differs from the one used in the previous section: the
915: solute-solvent interaction potential is purely repulsive and is given
916: by $U_i(r)=k_BT(r-R_0)^{-12}$, while the solute-solute interaction is
917: hard-sphere like with a hard sphere radius $R_0$. The solutes are
918: neutral or carry opposite charges $Q$ homogeneously distributed over
919: the sphere volume. The simulations were carried out with the SPC/E
920: model of water. In our theory, we fix the Tolman length to
921: $\delta=0.90$\AA~ as measured by Huang {\it et al.}\cite{huang:jpc}
922: and the dielectric constant to $\epsilon_l=71$ for SPC/E
923: water.\cite{berendsen:jpc} The mean forces are shown in
924: Fig.~\ref{fig:mf} for neutral spheres of radii $R_0=10$ and 12\AA~ and
925: oppositely charged solutes with radius $R_0=10$\AA~ and charge $Q=2$e
926: and 5e versus the solutes' surface-to-surface distance $s_0$.
927: Simulation and theory are in good, almost quantitative agreement, and
928: show that our theory captures the decreasing range of the strongly
929: hydrophobic attraction with decreasing radius and increasing
930: charge due to suppressed dewetting. We emphasize that our theory is
931: basically fit-parameter free for this system of large
932: solutes. Fig.~\ref{fig:mf} also shows the theoretical mean force for
933: the neutral $R_0=12$\AA~ solute using a smaller Tolman length
934: $\delta=0.75$\AA. The decrease of the Tolman length increases the
935: depth and range of the solvent-mediated solute-solute attractive mean
936: force by approximately 5\%, showing a nonvanishing but only slight
937: influence.
938: 
939: \section{Conclusion and final remarks}
940: 
941: In summary, we have presented a novel implicit solvent model which
942: couples polar and nonpolar solvation contributions by employing a
943: variational formalism in which the Gibbs free energy of the system is
944: expressed as a functional of the solvent volume exclusion
945: function. Minimization of the free energy leads to a Laplace-Young
946: like equation for the solvent excluded cavity around the solutes,
947: which is extended to describe solvation on mesoscopic and microscopic
948: scales.  We have shown that the theory gives a reasonable description
949: of the solvation of microscopic solutes, such as ions and
950: alkanes. Improved accuracy will require further refinement of the
951: curvature dependence of the surface tension $\gamma(\vec r;[v])$ and
952: the definition of the position-dependent dielectric constant
953: $\epsilon(\vec r;[v])$. Given the physically reasonable values of the
954: parameters $\delta$ and $\xi$ we found by fitting, we hope that
955: extensions based on physical rational, e.g. given by complementary
956: microscopic
957: approaches\cite{beglov:jcp,zwanzig,lum:jpc,hummer:pnas,garde:prl} and
958: further empirical corrections, will lead to an accurate fit-parameter
959: free implicit solvent description.
960: 
961: We have further demonstrated that on larger scales, where solvent
962: dewetting can play an important role in solvation, our formalism
963: captures the delicate balance between hydrophobic, dispersive and
964: electrostatic forces which has been observed in previous
965: systems.\cite{huang:jpcb,pettitt,dzubiella:channel1,vaitheesvaran,dzubiella:jcp:2003,zhou:science,berne:nature}
966: The dewetting in our model is driven by the interfacial term which
967: favors minimizing the solute-solvent interface. A comment must be made
968: here regarding the sensitivity of dewetting to the particular
969: solvent-solute interactions. As recently argued by
970: Chandler,\cite{chandler:review} extended fluid interfaces near phase
971: coexistence are often referred to as 'soft' because they can be
972: deformed with only little or no free-energy change.\cite{safran} Our
973: approach seems to account for this sensitivity since small changes of
974: the constraints in the differential equation (\ref{diff}) for the
975: shape function, given e.g. by the dispersion potential close to the
976: solute surface, can lead to a major deformation or even rupture
977: (wetting transition) of the inter-solute, dewetted region. As we have
978: shown, this can significantly change the pmf for the solutes. Thus we
979: anticipate that slight changes in the geometry of a system, e.g. a
980: slight concave or convex bending of two plate-like
981: solutes,\cite{huang:jpcb,pettitt} can lead to very different results
982: for the dewetting magnitude and the pmf.
983: 
984: The current illustrations utilized spherical and cylindrical
985: symmetries.  More complex molecules, such as proteins, will require
986: solving the full three dimensional problem.  Numerical algorithms for
987: the calculation of interface evolution for more complicated geometries
988: are provided by efficient level-set methods or fast marching
989: methods.\cite{level} We believe that in the full three-dimensional
990: (3D) case, our method will be much more efficient than other
991: microscopic approaches which partly resolve the water structure and
992: are able to describe dewetting effects, e.g. the Lum-Chandler-Weeks
993: theory\cite{lum:jpc} (LCW) or information theory
994: \cite{hummer:pnas,garde:prl} (IT), as only a two-dimensional surface
995: is sought rather than a 3D density distribution on a fine grid. We
996: remark that LCW and IT do not consider electrostatic interactions and
997: may benefit from our complementary approach.
998: 
999: \section*{Acknowledgment}
1000: 
1001: The authors thank Tushar Jain, John Mongan, and Cameron Mura for
1002: useful discussions.  J.D. and J.M.J.S acknowledge financial support
1003: from a DFG Forschungsstipendium and the PFC-sponsored Center for
1004: Theoretical Biological Physics (Grants No. PHY-0216576 and
1005: PHY-0225630), respectively. Work in the McCammon group is additionally
1006: supported by NIH, HHMI, NBCR, and Accelrys, Inc.
1007: 
1008: \section*{Appendix A: Curvatures in cylindrical coordinates}
1009: In our general parametrization for the shape function $r(z)$ we
1010: express the radial coordinate by $r=r(l)$ and the axial coordinate by
1011: $z=z(l)$, as functions of the parameter $l$. Depending on the geometry
1012: of the considered system, $l$ has to be conveniently chosen, for
1013: instance to be the arc length, or $r(l)=l$, or $z(l)=l$. In our
1014: illustration the most convenient choice is $z(l)=l$. The principal
1015: curvatures are generally given by\cite{frankel}
1016: \begin{eqnarray}
1017: \kappa_1(r,z)=\frac{-z'}{r\sqrt{z'^2+r'^2}} \;{\rm ,}\;\kappa_2(r,z)=\frac{z'r''-z''r'}{(z'^2+r'^2)^{3/2}},
1018: \end{eqnarray}
1019: where the primes indicate the partial derivative with respect to
1020: $l$. Additionally, the unit normal vector reads
1021: \begin{eqnarray}
1022: \vec
1023: n(r,z)=\frac{1}{\sqrt{z'^2+r'^2}}\left(\begin{array}{c}z'\\-r'\end{array}\right).
1024: \end{eqnarray}
1025: The differential equation (\ref{diff}) is then solved by a forward
1026: relaxation scheme in time $t$
1027: \begin{eqnarray}
1028: \left(\begin{array}{c}r(t+\Delta t)\\z(t+\Delta
1029: t)\end{array}\right)=\left(\begin{array}{c}r(t)\\z(t)\end{array}\right)-\Delta
1030: t\, {\vec n}(r,z){\rm de}(r,z),
1031: \end{eqnarray}
1032: where the steady-state solution $\partial(r,z)/\partial t=0$ is the
1033: solution of ${\rm de}(r,z)=0$ we are looking for.  In the numerical
1034: calculations we use a grid of 500 bins and an integration time step of
1035: $\Delta t=$0.001.  The first and second derivatives are approximated
1036: using a symmetric two and three-step finite difference equation,
1037: respectively. Convergence is usually reached after 10$^5$
1038: time steps. The result is observed to be independent of the initial
1039: choice of $r(z)$ at $t=0$.
1040: 
1041: 
1042: \section*{Appendix B: Numerical solution of the PB equation}
1043: Since we neglect mobile ions in our work, the PB equation reduces to
1044: Poisson's equation. It is solved on a two dimensional grid in
1045: cylindrical coordinates $r$ and $z$ with a finite difference method.
1046: The gradient and Laplacian are given then by $\nabla=(\partial r,
1047: \partial z)$ and $\Delta=\partial_r+\partial_r/r+\partial_z^2$,
1048: respectively. The first and second derivatives are approximated using
1049: symmetric two or three-step finite-difference equations. An explicit,
1050: forward time relaxation scheme is used to find the solution of
1051: Poisson's equation:
1052: \begin{eqnarray}
1053: \Psi(t+\Delta t;\vec r)=\Psi(t;\vec r)-\Delta t {\rm PB}(\Psi(t;\vec r)).
1054: \end{eqnarray}
1055:  In most cases we use a lattice spacing of $\Delta r=\Delta z=0.4$\AA~
1056: on a $n_r\times n_z=100\times 200$ grid, and an integration time step
1057: $\Delta t=0.05$. Convergence takes approximately $10^5$ time
1058: steps. For the charged particles which are buried in the nanosolutes
1059: we use homogeneously charged spheres with a radius of 2\AA.  Instead
1060: of a sharp transition for the dielectric boundary ~(\ref{e}), we use a
1061: smoothing function for reasons of numerical stability:
1062: \begin{eqnarray}
1063: \epsilon(\vec r)=\frac{\epsilon_l-\epsilon_v}{\exp(\kappa d(\vec r))+1}+\epsilon_v,
1064: \label{e3}
1065: \end{eqnarray}
1066: where the absolute value of the length $d(\vec r)$ is given by the nearest
1067: distance to the boundary of the volume exclusion function $v(\vec
1068: r)$. $d$ is defined to be positive when $\vec r\in \cal V$
1069: and negative elsewhere. The inverse length $\kappa$ defines the width
1070: of the boundary region and in the limit $\kappa\rightarrow \infty$ we
1071: recover the sharp transition (\ref{e}).  We choose a value $\kappa
1072: \gtrsim 3{\rm \AA}^{-1}$ for which the solution of Poisson's equation
1073: becomes basically independent of the choice of $\kappa$. An example for the
1074: dielectric boundary is shown in Fig.~\ref{eps} for two partly dewetted
1075: nanosolutes of radius $R_0=15$\AA~ at a distance $s_0=7$\AA~ carrying
1076: a charge $Q=5e$ (system V in sec. III.C). 
1077: 
1078: In order to obtain the optimal shape function $v_{\rm min}(\vec r)$
1079: the shape equation (\ref{diff}) has to be solved simultaneously with
1080: Poisson's equation when the solutes are charged. In practice, we
1081: first solve (\ref{diff}) without any electrostatic contributions. In
1082: the second step, we solve  Poisson's equation with the dielectric
1083: boundary (\ref{e3}) given by the volume exclusion function of the
1084: former solution. The result for the electric energy density is then
1085: plugged back in the shape equation in the third step. The last two
1086: steps are repeated until the solution for $v_{\rm min}(\vec r)$ is
1087: fully converged. Since the results for $r(z)$ excluding and including
1088: electrostatics are quite similar for our systems, full convergence
1089: takes usually only 6 to 7 repetitions of the described iteration
1090: steps.
1091: 
1092: \begin{thebibliography}{71}
1093: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1094: \expandafter\ifx\csname bibnamefont\endcsname\relax
1095:   \def\bibnamefont#1{#1}\fi
1096: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1097:   \def\bibfnamefont#1{#1}\fi
1098: \expandafter\ifx\csname citenamefont\endcsname\relax
1099:   \def\citenamefont#1{#1}\fi
1100: \expandafter\ifx\csname url\endcsname\relax
1101:   \def\url#1{\texttt{#1}}\fi
1102: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1103: \providecommand{\bibinfo}[2]{#2}
1104: \providecommand{\eprint}[2][]{\url{#2}}
1105: 
1106: \bibitem[{\citenamefont{Roux and Simonson}(1999)}]{roux:biochem}
1107: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Roux}} \bibnamefont{and}
1108:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Simonson}},
1109:   \bibinfo{journal}{Biophys. Chem.} \textbf{\bibinfo{volume}{78}},
1110:   \bibinfo{pages}{1} (\bibinfo{year}{1999}).
1111: 
1112: \bibitem[{\citenamefont{Richards}(1977)}]{richards}
1113: \bibinfo{author}{\bibfnamefont{F.~M.} \bibnamefont{Richards}},
1114:   \bibinfo{journal}{Annu. Rev. Biophys. Bioeng.} \textbf{\bibinfo{volume}{6}},
1115:   \bibinfo{pages}{151} (\bibinfo{year}{1977}).
1116: 
1117: \bibitem[{\citenamefont{Gallicchio et~al.}(2000)\citenamefont{Gallicchio, Kubo,
1118:   and Levy}}]{gallicchio:jpcb}
1119: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gallicchio}},
1120:   \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Kubo}}, \bibnamefont{and}
1121:   \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{Levy}},
1122:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{104}},
1123:   \bibinfo{pages}{6271} (\bibinfo{year}{2000}).
1124: 
1125: \bibitem[{\citenamefont{Gallicchio et~al.}(2002)\citenamefont{Gallicchio,
1126:   Zhang, and Levy}}]{gallicchio:jcc}
1127: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gallicchio}},
1128:   \bibinfo{author}{\bibfnamefont{L.~Y.} \bibnamefont{Zhang}}, \bibnamefont{and}
1129:   \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{Levy}},
1130:   \bibinfo{journal}{J. Comput. Chem.} \textbf{\bibinfo{volume}{23}},
1131:   \bibinfo{pages}{517} (\bibinfo{year}{2002}).
1132: 
1133: \bibitem[{\citenamefont{Zacharias}(2003)}]{zacharias}
1134: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zacharias}},
1135:   \bibinfo{journal}{J. Chem. Phys. A} \textbf{\bibinfo{volume}{107}},
1136:   \bibinfo{pages}{3000} (\bibinfo{year}{2003}).
1137: 
1138: \bibitem[{\citenamefont{Su and Gallicchio}(2004)}]{su:biochem}
1139: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Su}} \bibnamefont{and}
1140:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gallicchio}},
1141:   \bibinfo{journal}{Biophys. Chem.} \textbf{\bibinfo{volume}{109}},
1142:   \bibinfo{pages}{251} (\bibinfo{year}{2004}).
1143: 
1144: \bibitem[{\citenamefont{Levy et~al.}(2003)\citenamefont{Levy, Zhang,
1145:   Gallicchio, and Felts}}]{levy:jacs}
1146: \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{Levy}},
1147:   \bibinfo{author}{\bibfnamefont{L.~Y.} \bibnamefont{Zhang}},
1148:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gallicchio}},
1149:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~K.} \bibnamefont{Felts}},
1150:   \bibinfo{journal}{J. Am. Chem. Soc.} \textbf{\bibinfo{volume}{125}},
1151:   \bibinfo{pages}{9523} (\bibinfo{year}{2003}).
1152: 
1153: \bibitem[{\citenamefont{Bashford and Case}(2000)}]{bashford}
1154: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bashford}} \bibnamefont{and}
1155:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Case}},
1156:   \bibinfo{journal}{Ann. Rev. Phys. Chem} \textbf{\bibinfo{volume}{51}},
1157:   \bibinfo{pages}{129} (\bibinfo{year}{2000}).
1158: 
1159: \bibitem[{\citenamefont{Sharp and Honig}(1990)}]{sharp:honig}
1160: \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Sharp}} \bibnamefont{and}
1161:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Honig}}, \bibinfo{journal}{J.
1162:   Phys. Chem.} \textbf{\bibinfo{volume}{94}}, \bibinfo{pages}{7684}
1163:   (\bibinfo{year}{1990}).
1164: 
1165: \bibitem[{\citenamefont{Baker}(2005)}]{baker}
1166: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Baker}},
1167:   \bibinfo{journal}{Curr. Opin. Struct. Bio} \textbf{\bibinfo{volume}{15}},
1168:   \bibinfo{pages}{137} (\bibinfo{year}{2005}).
1169: 
1170: \bibitem[{\citenamefont{Nina et~al.}(1997)\citenamefont{Nina, Beglov, and
1171:   Roux}}]{nina}
1172: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nina}},
1173:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Beglov}}, \bibnamefont{and}
1174:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Roux}}, \bibinfo{journal}{J.
1175:   Phys. Chem. B} \textbf{\bibinfo{volume}{101}}, \bibinfo{pages}{5239}
1176:   (\bibinfo{year}{1997}).
1177: 
1178: \bibitem[{\citenamefont{Huang and Chandler}(2002)}]{huang:jpcb:2002}
1179: \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{Huang}} \bibnamefont{and}
1180:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}},
1181:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{106}},
1182:   \bibinfo{pages}{2047} (\bibinfo{year}{2002}).
1183: 
1184: \bibitem[{\citenamefont{Ashbaugh et~al.}(1998)\citenamefont{Ashbaugh, Kaler,
1185:   and Paulaitis}}]{ashbaugh:biophys}
1186: \bibinfo{author}{\bibfnamefont{H.~S.} \bibnamefont{Ashbaugh}},
1187:   \bibinfo{author}{\bibfnamefont{E.~W.} \bibnamefont{Kaler}}, \bibnamefont{and}
1188:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Paulaitis}},
1189:   \bibinfo{journal}{Biophys. J.} \textbf{\bibinfo{volume}{75}},
1190:   \bibinfo{pages}{755} (\bibinfo{year}{1998}).
1191: 
1192: \bibitem[{\citenamefont{Stillinger}(1973)}]{stilinger}
1193: \bibinfo{author}{\bibfnamefont{F.~H.} \bibnamefont{Stillinger}},
1194:   \bibinfo{journal}{J. Solution Chem.} \textbf{\bibinfo{volume}{2}},
1195:   \bibinfo{pages}{141} (\bibinfo{year}{1973}).
1196: 
1197: \bibitem[{\citenamefont{Lum et~al.}(1999)\citenamefont{Lum, Chandler, and
1198:   Weeks}}]{lum:jpc}
1199: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lum}},
1200:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}}, \bibnamefont{and}
1201:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Weeks}},
1202:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{103}},
1203:   \bibinfo{pages}{4570} (\bibinfo{year}{1999}).
1204: 
1205: \bibitem[{\citenamefont{Hummer et~al.}(1997)\citenamefont{Hummer, Pratt, and
1206:   Garcia}}]{hummer:jcp:1997}
1207: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1208:   \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{Pratt}}, \bibnamefont{and}
1209:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1210:   \bibinfo{journal}{J. Chem. Phys} \textbf{\bibinfo{volume}{107}},
1211:   \bibinfo{pages}{9275} (\bibinfo{year}{1997}).
1212: 
1213: \bibitem[{\citenamefont{Chandler}(2005)}]{chandler:review}
1214: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}},
1215:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{437}},
1216:   \bibinfo{pages}{640} (\bibinfo{year}{2005}).
1217: 
1218: \bibitem[{\citenamefont{Wallquist and Berne}(1995)}]{wallquist:jpc}
1219: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Wallquist}} \bibnamefont{and}
1220:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1221:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{99}},
1222:   \bibinfo{pages}{2893} (\bibinfo{year}{1995}).
1223: 
1224: \bibitem[{\citenamefont{Huang et~al.}(2003)\citenamefont{Huang, Margulis, and
1225:   Berne}}]{huang:pnas}
1226: \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Huang}},
1227:   \bibinfo{author}{\bibfnamefont{C.~J.} \bibnamefont{Margulis}},
1228:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1229:   \bibinfo{journal}{PNAS} \textbf{\bibinfo{volume}{100}},
1230:   \bibinfo{pages}{11953} (\bibinfo{year}{2003}).
1231: 
1232: \bibitem[{\citenamefont{Huang et~al.}(2005)\citenamefont{Huang, Zhou, and
1233:   Berne}}]{huang:jpcb}
1234: \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Huang}},
1235:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zhou}}, \bibnamefont{and}
1236:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1237:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{109}},
1238:   \bibinfo{pages}{3546} (\bibinfo{year}{2005}).
1239: 
1240: \bibitem[{\citenamefont{Choudhury and Pettitt}(2005)}]{pettitt}
1241: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Choudhury}} \bibnamefont{and}
1242:   \bibinfo{author}{\bibfnamefont{B.~M.} \bibnamefont{Pettitt}},
1243:   \bibinfo{journal}{J. Am. Chem. Soc.} \textbf{\bibinfo{volume}{127}},
1244:   \bibinfo{pages}{3556} (\bibinfo{year}{2005}).
1245: 
1246: \bibitem[{\citenamefont{Hummer et~al.}(2001)\citenamefont{Hummer, Rasaiah, and
1247:   Nowortya}}]{hummer:nature}
1248: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1249:   \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Rasaiah}},
1250:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~P.}
1251:   \bibnamefont{Nowortya}}, \bibinfo{journal}{Nature}
1252:   \textbf{\bibinfo{volume}{414}}, \bibinfo{pages}{188} (\bibinfo{year}{2001}).
1253: 
1254: \bibitem[{\citenamefont{Allen et~al.}(2002)\citenamefont{Allen, Melchionna, and
1255:   Hansen}}]{allen:prl}
1256: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Allen}},
1257:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Melchionna}},
1258:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.-P.}
1259:   \bibnamefont{Hansen}}, \bibinfo{journal}{Phys. Rev. Lett.}
1260:   \textbf{\bibinfo{volume}{89}}, \bibinfo{pages}{175502}
1261:   (\bibinfo{year}{2002}).
1262: 
1263: \bibitem[{\citenamefont{Beckstein et~al.}(2001)\citenamefont{Beckstein, Biggin,
1264:   and Sansom}}]{beckstein:jpc}
1265: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Beckstein}},
1266:   \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Biggin}},
1267:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~S.~P.}
1268:   \bibnamefont{Sansom}}, \bibinfo{journal}{J. Phys. Chem.}
1269:   \textbf{\bibinfo{volume}{105}}, \bibinfo{pages}{12902}
1270:   (\bibinfo{year}{2001}).
1271: 
1272: \bibitem[{\citenamefont{Anishkin and Sukharev}(2004)}]{sukharev}
1273: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Anishkin}} \bibnamefont{and}
1274:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sukharev}},
1275:   \bibinfo{journal}{Biophys. J.} \textbf{\bibinfo{volume}{86}},
1276:   \bibinfo{pages}{2883} (\bibinfo{year}{2004}).
1277: 
1278: \bibitem[{\citenamefont{Dzubiella and
1279:   Hansen}(2004{\natexlab{a}})}]{dzubiella:channel1}
1280: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}} \bibnamefont{and}
1281:   \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1282:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{120}},
1283:   \bibinfo{pages}{5001} (\bibinfo{year}{2004}{\natexlab{a}}).
1284: 
1285: \bibitem[{\citenamefont{Vaitheesvaran et~al.}(2004)\citenamefont{Vaitheesvaran,
1286:   Rasaiah, and Hummer}}]{vaitheesvaran}
1287: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Vaitheesvaran}},
1288:   \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Rasaiah}},
1289:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1290:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{121}},
1291:   \bibinfo{pages}{7955} (\bibinfo{year}{2004}).
1292: 
1293: \bibitem[{\citenamefont{Dzubiella and Hansen}(2005)}]{dzubiella:channel2}
1294: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}} \bibnamefont{and}
1295:   \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1296:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{122}},
1297:   \bibinfo{pages}{234706} (\bibinfo{year}{2005}).
1298: 
1299: \bibitem[{\citenamefont{Dzubiella and Hansen}(2003)}]{dzubiella:jcp:2003}
1300: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}} \bibnamefont{and}
1301:   \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1302:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{119}},
1303:   \bibinfo{pages}{12049} (\bibinfo{year}{2003}).
1304: 
1305: \bibitem[{\citenamefont{Dzubiella and
1306:   Hansen}(2004{\natexlab{b}})}]{dzubiella:jcp:2004}
1307: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}} \bibnamefont{and}
1308:   \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1309:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{121}},
1310:   \bibinfo{pages}{5514} (\bibinfo{year}{2004}{\natexlab{b}}).
1311: 
1312: \bibitem[{\citenamefont{Zhou et~al.}(2004)\citenamefont{Zhou, Huang, Margulis,
1313:   and Berne}}]{zhou:science}
1314: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zhou}},
1315:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Huang}},
1316:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Margulis}}, \bibnamefont{and}
1317:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1318:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{305}},
1319:   \bibinfo{pages}{1605} (\bibinfo{year}{2004}).
1320: 
1321: \bibitem[{\citenamefont{Liu et~al.}(2005)\citenamefont{Liu, Huang, Zhou, and
1322:   Berne}}]{berne:nature}
1323: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Liu}},
1324:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Huang}},
1325:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zhou}}, \bibnamefont{and}
1326:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1327:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{437}},
1328:   \bibinfo{pages}{159} (\bibinfo{year}{2005}).
1329: 
1330: \bibitem[{\citenamefont{Parker et~al.}(1994)\citenamefont{Parker, Claesson, and
1331:   Attard}}]{attard}
1332: \bibinfo{author}{\bibfnamefont{J.~L.} \bibnamefont{Parker}},
1333:   \bibinfo{author}{\bibfnamefont{P.~M.} \bibnamefont{Claesson}},
1334:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Attard}},
1335:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{98}},
1336:   \bibinfo{pages}{8468} (\bibinfo{year}{1994}).
1337: 
1338: \bibitem[{\citenamefont{Beglov and Roux}(1996)}]{beglov:jcp}
1339: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Beglov}} \bibnamefont{and}
1340:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Roux}}, \bibinfo{journal}{J.
1341:   Chem. Phys.} \textbf{\bibinfo{volume}{104}}, \bibinfo{pages}{8678}
1342:   (\bibinfo{year}{1996}).
1343: 
1344: \bibitem[{\citenamefont{Kralchevsky and Nagayama}(Amsterdam)}]{kralchevsky}
1345: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Kralchevsky}} \bibnamefont{and}
1346:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Nagayama}},
1347:   \emph{\bibinfo{title}{Particles at Fluid Interfaces and Membranes}}
1348:   (\bibinfo{publisher}{Elsevier}, \bibinfo{address}{2001},
1349:   \bibinfo{year}{Amsterdam}).
1350: 
1351: \bibitem[{\citenamefont{Helfrich}(1973)}]{helfrich}
1352: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Helfrich}},
1353:   \bibinfo{journal}{Z. Naturforsch. C} \textbf{\bibinfo{volume}{28}},
1354:   \bibinfo{pages}{693} (\bibinfo{year}{1973}).
1355: 
1356: \bibitem[{\citenamefont{Zhong-can and Helfrich}(1989)}]{helfrich2}
1357: \bibinfo{author}{\bibfnamefont{O.-Y.} \bibnamefont{Zhong-can}}
1358:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Helfrich}},
1359:   \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{39}},
1360:   \bibinfo{pages}{5280} (\bibinfo{year}{1989}).
1361: 
1362: \bibitem[{\citenamefont{Bieker and Dietrich}(1998)}]{bieker}
1363: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Bieker}} \bibnamefont{and}
1364:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1365:   \bibinfo{journal}{Physica A} \textbf{\bibinfo{volume}{252}},
1366:   \bibinfo{pages}{85} (\bibinfo{year}{1998}).
1367: 
1368: \bibitem[{\citenamefont{Chou}(2001)}]{electrowetting}
1369: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Chou}}, \bibinfo{journal}{Phys.
1370:   Rev. Lett.} \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{106101}
1371:   (\bibinfo{year}{2001}).
1372: 
1373: \bibitem[{\citenamefont{Allen et~al.}(2003)\citenamefont{Allen, Melchionna, and
1374:   Hansen}}]{allen:jcp}
1375: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Allen}},
1376:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Melchionna}},
1377:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.-P.}
1378:   \bibnamefont{Hansen}}, \bibinfo{journal}{J. Chem. Phys.}
1379:   \textbf{\bibinfo{volume}{119}}, \bibinfo{pages}{3905} (\bibinfo{year}{2003}).
1380: 
1381: \bibitem[{\citenamefont{Attard}(2000)}]{attard2}
1382: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Attard}},
1383:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{16}},
1384:   \bibinfo{pages}{4455} (\bibinfo{year}{2000}).
1385: 
1386: \bibitem[{\citenamefont{Cheng and Rossky}(1998)}]{rossky:nature}
1387: \bibinfo{author}{\bibfnamefont{Y.-K.} \bibnamefont{Cheng}} \bibnamefont{and}
1388:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Rossky}},
1389:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{392}},
1390:   \bibinfo{pages}{696} (\bibinfo{year}{1998}).
1391: 
1392: \bibitem[{\citenamefont{Gerstein and Chothia}(1996)}]{gerstein}
1393: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Gerstein}} \bibnamefont{and}
1394:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Chothia}},
1395:   \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{93}},
1396:   \bibinfo{pages}{10167} (\bibinfo{year}{1996}).
1397: 
1398: \bibitem[{\citenamefont{Lau et~al.}(2003)\citenamefont{Lau, Bico, Teo,
1399:   Chhowalla, Amaratunga, Milne, McKinley, and Gleason}}]{super}
1400: \bibinfo{author}{\bibfnamefont{K.~K.~S.} \bibnamefont{Lau}},
1401:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bico}},
1402:   \bibinfo{author}{\bibfnamefont{K.~B.~K.} \bibnamefont{Teo}},
1403:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Chhowalla}},
1404:   \bibinfo{author}{\bibfnamefont{G.~A.~J.} \bibnamefont{Amaratunga}},
1405:   \bibinfo{author}{\bibfnamefont{W.~I.} \bibnamefont{Milne}},
1406:   \bibinfo{author}{\bibfnamefont{G.~H.} \bibnamefont{McKinley}},
1407:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.~K.}
1408:   \bibnamefont{Gleason}}, \bibinfo{journal}{Nano Letters}
1409:   \textbf{\bibinfo{volume}{3}}, \bibinfo{pages}{1701} (\bibinfo{year}{2003}).
1410: 
1411: \bibitem[{\citenamefont{Dzubiella et~al.}(2005)\citenamefont{Dzubiella,
1412:   Swanson, and McCammon}}]{prl}
1413: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}},
1414:   \bibinfo{author}{\bibfnamefont{J.~M.~J.} \bibnamefont{Swanson}},
1415:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~A.}
1416:   \bibnamefont{McCammon}} (\bibinfo{year}{2005}), \bibinfo{note}{to be
1417:   published}.
1418: 
1419: \bibitem[{\citenamefont{Triezenberg and Zwanzig}(1972)}]{zwanzig}
1420: \bibinfo{author}{\bibfnamefont{D.~G.} \bibnamefont{Triezenberg}}
1421:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zwanzig}},
1422:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{28}},
1423:   \bibinfo{pages}{1183} (\bibinfo{year}{1972}).
1424: 
1425: \bibitem[{\citenamefont{Jackson}(1999)}]{jackson}
1426: \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Jackson}},
1427:   \emph{\bibinfo{title}{Classical Electrodynamics}}
1428:   (\bibinfo{publisher}{Wiley}, \bibinfo{address}{New York},
1429:   \bibinfo{year}{1999}), \bibinfo{note}{3rd edn.}
1430: 
1431: \bibitem[{\citenamefont{Gilson et~al.}(1993)\citenamefont{Gilson, Davis, Luty,
1432:   and McCammon}}]{gilson}
1433: \bibinfo{author}{\bibfnamefont{M.~K.} \bibnamefont{Gilson}},
1434:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Davis}},
1435:   \bibinfo{author}{\bibfnamefont{B.~A.} \bibnamefont{Luty}}, \bibnamefont{and}
1436:   \bibinfo{author}{\bibfnamefont{J.~A.} \bibnamefont{McCammon}},
1437:   \bibinfo{journal}{J. Phys. Chem} \textbf{\bibinfo{volume}{97}},
1438:   \bibinfo{pages}{3591} (\bibinfo{year}{1993}).
1439: 
1440: \bibitem[{\citenamefont{Huang et~al.}(2001)\citenamefont{Huang, Geissler, and
1441:   Chandler}}]{huang:jpc}
1442: \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{Huang}},
1443:   \bibinfo{author}{\bibfnamefont{P.~L.} \bibnamefont{Geissler}},
1444:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}},
1445:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{105}},
1446:   \bibinfo{pages}{6704} (\bibinfo{year}{2001}).
1447: 
1448: \bibitem[{pri()}]{prince}
1449: \bibinfo{note}{The principal curvatures are formally given by the eigenvectors
1450:   of the Hessian (or shape operator) of $v$, $\hat S$, which can be expressed
1451:   by the vector gradient of the unit normal vector field $\vec n(\vec r)$ of
1452:   the surface, $\hat S=\nabla \vec n(\vec r)={\bf \nabla} \frac{\nabla v(\vec
1453:   r)}{|\nabla v(\vec r)|}$.}
1454: 
1455: \bibitem[{\citenamefont{Henderson}(1992)}]{hendersonreview}
1456: \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Henderson}},
1457:   \emph{\bibinfo{title}{Fundamentals of Inhomogeneous Fluids}}
1458:   (\bibinfo{publisher}{Dekker}, \bibinfo{address}{New York},
1459:   \bibinfo{year}{1992}), \bibinfo{note}{edited by D.~Henderson}.
1460: 
1461: \bibitem[{\citenamefont{Stewart and Evans}(2005)}]{evans2}
1462: \bibinfo{author}{\bibfnamefont{M.~C.} \bibnamefont{Stewart}} \bibnamefont{and}
1463:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Evans}},
1464:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{71}},
1465:   \bibinfo{pages}{011602} (\bibinfo{year}{2005}).
1466: 
1467: \bibitem[{\citenamefont{Reiss}(1965)}]{spt}
1468: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Reiss}}, \bibinfo{journal}{Adv.
1469:   Chem. Phys.} \textbf{\bibinfo{volume}{9}}, \bibinfo{pages}{1}
1470:   (\bibinfo{year}{1965}).
1471: 
1472: \bibitem[{\citenamefont{Tolman}(1949)}]{tolman}
1473: \bibinfo{author}{\bibfnamefont{R.~C.} \bibnamefont{Tolman}},
1474:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{17}},
1475:   \bibinfo{pages}{333} (\bibinfo{year}{1949}).
1476: 
1477: \bibitem[{\citenamefont{Nicholls et~al.}(1991)\citenamefont{Nicholls, Sharp,
1478:   and Honig}}]{nicholls}
1479: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Nicholls}},
1480:   \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Sharp}}, \bibnamefont{and}
1481:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Honig}},
1482:   \bibinfo{journal}{Proteins} \textbf{\bibinfo{volume}{11}},
1483:   \bibinfo{pages}{281} (\bibinfo{year}{1991}).
1484: 
1485: \bibitem[{\citenamefont{Luo and Tucker}(1996)}]{luo:jpc}
1486: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Luo}} \bibnamefont{and}
1487:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Tucker}},
1488:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{100}},
1489:   \bibinfo{pages}{11165} (\bibinfo{year}{1996}).
1490: 
1491: \bibitem[{\citenamefont{Berendsen et~al.}(1987)\citenamefont{Berendsen,
1492:   Grigera, and Straatsma}}]{berendsen:jpc}
1493: \bibinfo{author}{\bibfnamefont{H.~J.~C.} \bibnamefont{Berendsen}},
1494:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Grigera}},
1495:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.~P.}
1496:   \bibnamefont{Straatsma}}, \bibinfo{journal}{J. Phys. Chem.}
1497:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{6269} (\bibinfo{year}{1987}).
1498: 
1499: \bibitem[{\citenamefont{Manjari and Kim}(2005)}]{kim}
1500: \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{Manjari}} \bibnamefont{and}
1501:   \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Kim}},
1502:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{123}},
1503:   \bibinfo{pages}{014504} (\bibinfo{year}{2005}).
1504: 
1505: \bibitem[{\citenamefont{Hummer et~al.}(1996{\natexlab{a}})\citenamefont{Hummer,
1506:   Pratt, and Garcia}}]{hummer:jpc:1996}
1507: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1508:   \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{Pratt}}, \bibnamefont{and}
1509:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1510:   \bibinfo{journal}{J. Phys. Chem} \textbf{\bibinfo{volume}{100}},
1511:   \bibinfo{pages}{1206} (\bibinfo{year}{1996}{\natexlab{a}}).
1512: 
1513: \bibitem[{\citenamefont{Paschek}(2004)}]{paschek}
1514: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Paschek}}, \bibinfo{journal}{J.
1515:   Chem. Phys.} \textbf{\bibinfo{volume}{120}}, \bibinfo{pages}{6674}
1516:   (\bibinfo{year}{2004}).
1517: 
1518: \bibitem[{\citenamefont{Alejandre et~al.}(1995)\citenamefont{Alejandre,
1519:   Tildesley, and Chapela}}]{alejandre}
1520: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Alejandre}},
1521:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Tildesley}},
1522:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~A.}
1523:   \bibnamefont{Chapela}}, \bibinfo{journal}{J. Chem. Phys.}
1524:   \textbf{\bibinfo{volume}{102}}, \bibinfo{pages}{4574} (\bibinfo{year}{1995}).
1525: 
1526: \bibitem[{\citenamefont{Connolly}(1993)}]{connolly}
1527: \bibinfo{author}{\bibfnamefont{M.~L.} \bibnamefont{Connolly}},
1528:   \bibinfo{journal}{J. Mol. Graphics} \textbf{\bibinfo{volume}{11}},
1529:   \bibinfo{pages}{139} (\bibinfo{year}{1993}).
1530: 
1531: \bibitem[{sig()}]{sigmanote}
1532: \bibinfo{note}{The solute-solute LJ parameters have been calculated from the
1533:   solute-water LJ parameters employing the usual combining rules. Note that the
1534:   rules used by Ashbaugh et al.,\cite{ashbaugh:biophys} are different to those
1535:   of Hummer et al.\cite{hummer:jpc:1996} and Paschek\cite{paschek}}.
1536: 
1537: \bibitem[{\citenamefont{van~der Spoel et~al.}(1998)\citenamefont{van~der Spoel,
1538:   van Maaren, and Berendsen}}]{spoel}
1539: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{van~der Spoel}},
1540:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{van Maaren}},
1541:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~J.~C.}
1542:   \bibnamefont{Berendsen}}, \bibinfo{journal}{J. Chem. Phys.}
1543:   \textbf{\bibinfo{volume}{108}}, \bibinfo{pages}{10220}
1544:   (\bibinfo{year}{1998}).
1545: 
1546: \bibitem[{\citenamefont{Jorgensen et~al.}(1984)\citenamefont{Jorgensen, Madura,
1547:   and Swenson}}]{jorgensen}
1548: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Jorgensen}},
1549:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Madura}},
1550:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.~J.}
1551:   \bibnamefont{Swenson}}, \bibinfo{journal}{J. Am. Chem. Soc.}
1552:   \textbf{\bibinfo{volume}{106}}, \bibinfo{pages}{6638} (\bibinfo{year}{1984}).
1553: 
1554: \bibitem[{\citenamefont{Fernandez}(2004)}]{fernandez}
1555: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fernandez}},
1556:   \bibinfo{journal}{Nature Biotech.} \textbf{\bibinfo{volume}{22}},
1557:   \bibinfo{pages}{1081} (\bibinfo{year}{2004}).
1558: 
1559: \bibitem[{\citenamefont{Hummer et~al.}(1996{\natexlab{b}})\citenamefont{Hummer,
1560:   Garde, Garcia, Phorille, and Pratt}}]{hummer:pnas}
1561: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1562:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Garde}},
1563:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1564:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Phorille}}, \bibnamefont{and}
1565:   \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{Pratt}},
1566:   \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{93}},
1567:   \bibinfo{pages}{8951} (\bibinfo{year}{1996}{\natexlab{b}}).
1568: 
1569: \bibitem[{\citenamefont{Garde et~al.}(1996)\citenamefont{Garde, Hummer, Garcia,
1570:   Paulaitis, and Pratt}}]{garde:prl}
1571: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Garde}},
1572:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1573:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1574:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Paulaitis}},
1575:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{Pratt}},
1576:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{77}},
1577:   \bibinfo{pages}{4966} (\bibinfo{year}{1996}).
1578: 
1579: \bibitem[{\citenamefont{Safran}(1994)}]{safran}
1580: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Safran}},
1581:   \emph{\bibinfo{title}{Statistical Thermodynamics of Surfaces, Interfaces and
1582:   Membranes}} (\bibinfo{publisher}{Addison-Wesley}, \bibinfo{address}{Reading},
1583:   \bibinfo{year}{1994}), \bibinfo{note}{ch. 1-3}.
1584: 
1585: \bibitem[{\citenamefont{Sethian}(1999)}]{level}
1586: \bibinfo{author}{\bibfnamefont{J.~A.} \bibnamefont{Sethian}},
1587:   \emph{\bibinfo{title}{Level Set methods and Fast Marching Methods: Evolving
1588:   Interfaces in Geometry, Fluid Mechanics, Computer Vision, and Materials
1589:   Science}} (\bibinfo{publisher}{Cambridge University Press},
1590:   \bibinfo{address}{Cambridge}, \bibinfo{year}{1999}).
1591: 
1592: \bibitem[{\citenamefont{Frankel}(1997)}]{frankel}
1593: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Frankel}},
1594:   \emph{\bibinfo{title}{The Geometry of Physics}}
1595:   (\bibinfo{publisher}{Cambridge University Press},
1596:   \bibinfo{address}{Cambridge}, \bibinfo{year}{1997}).
1597: 
1598: \end{thebibliography}
1599: 
1600: \newpage
1601: 
1602: \begin{table}
1603: \begin{center}
1604: \begin{tabular}{l | c c c | c c | c }
1605: Solute & $\epsilon/({\rm kJ/mol})$\; & $\sigma/$\AA\; &
1606: $\Delta G_{\rm sim}/({\rm kJ/mol})$ & $\delta_{\rm bf}/$\AA &
1607: $R_{\rm min}/$\AA \\  \hline \hline SPC & 0.65 & 3.17 & -- & -- & --
1608: \\ SPC/E & 0.65 & 3.17 & -- & -- & -- \\ \hline \hline
1609: ref\cite{hummer:jpc:1996} & & & & & \\  Na$^0$ & 0.2005 & 2.85 &
1610: 9.2(1) & 0.79 & 2.32 \\ K$^0$ & 0.0061 & 4.52 & 23.7(5) & 0.76 & 2.83
1611: \\ Ca$^{0}$ & 0.6380 & 3.17 & 10.2(3) & 0.85 & 2.80 \\ F$^0$ & 0.5538
1612: & 3.05 & 9.7(2) & 0.85 & 2.68 \\ Cl$^0$ & 0.5380 & 3.75 & 21(3) & 0.80
1613: & 3.30 \\  Br$^0$ & 0.4945 & 3.83 & 24(3) & 0.77 & 3.35 \\  \hline
1614: \hline  ref\cite{paschek} & & & & & \\  Ne & 0.3156 & 3.10 &
1615: 11.41(0.05)& 0.84 & 2.61 \\  Ar & 0.8176 & 3.29 & 8.68 (0.08) & 0.90 &
1616: 2.96\\  Kr & 0.9518 & 3.42 & 8.12 (0.1) & 0.91 & 3.10 \\  Xe & 1.0710
1617: & 3.57 & 7.65 (0.15) & 0.92 & 3.27 \\ \hline  \hline
1618: ref\cite{ashbaugh:biophys} & & & & & \\ 
1619: Me & 0.8941 &3.44 & 10.96 (0.46) & 0.85 & 3.11 \\  
1620: \hline 
1621: ethane & -- & -- & 10.75 (0.50) & 0.87 & -- \\
1622: CH$_3$ & 0.7503 & 3.46&-- & -- & -- \\  
1623: \hline 
1624: propane & -- & -- & 13.81 (0.54) & 0.94 &-- \\ 
1625: butane & -- & -- & 14.69 (0.54) & 0.96 &-- \\ 
1626: CH$_2$ & 0.5665& 3.52 & -- & -- & -- \\ 
1627: CH$_3$ & 0.6900 & 3.52 & -- & -- & -- \\
1628: \end{tabular}
1629: \caption{Solute-water LJ parameters and solvation free energy $\Delta
1630:  G_{\rm sim}$ for neutral Lennard-Jones spheres from the SPC water
1631:  simulations performed by Hummer {\it et al.}\cite{hummer:jpc:1996}
1632:  and Paschek.\cite{paschek} $\delta_{\rm bf}$ is the Tolman length
1633:  best fit to $\Delta G_{\rm sim}$ (rounded to two digits after the
1634:  decimal point). $R_{\rm min}$ is the resulting optimal radius
1635:  excluded of solvent. Also shown are the values for the simple alkanes
1636:  methane (Me), ethane, propane, and butane from the study of Ashbaugh
1637:  {\it et al.}\cite{ashbaugh:biophys} Simulation errors are given in
1638:  parentheses.}
1639: \label{tab1}
1640: \end{center}
1641: \end{table} 
1642: 
1643: 
1644: \begin{table}
1645: \begin{center}
1646: \begin{tabular}{l | c c c}
1647:    Atom & $\Delta G_{\rm sim}/({\rm kJ\,mol}^{-1})$ & $\delta_{\rm
1648: bf}/$\AA & $R_{\rm min}/$\AA \\  \hline Ne & 11.65 (0.05) &0.88 & 2.60
1649: \\ Ar & 8.83 (0.08) &0.96 & 2.94 \\ Kr & 8.20 (0.1) &0.98 & 3.09 \\ Xe
1650: & 7.58 (0.15) &1.00 & 3.25 \\
1651: \end{tabular}
1652: \caption{Solvation free energies for neutral Lennard-Jones spheres in
1653: SPC/E water from the simulations of Paschek.\cite{paschek}
1654: $\delta_{\rm bf}$ and $R_{\rm min}$ are defined as in Tab.~I.}
1655: \label{tab1}
1656: \end{center}
1657: \end{table} 
1658: 
1659: \begin{table}
1660: \begin{center}
1661: \begin{tabular}{l | c c c c c}
1662:   Ion & $q$ & $\frac{\Delta G_{\rm sim}}{{\rm kJ\,mol}^{-1}}$ & $\frac{\Delta G}{{\rm kJ\,mol}^{-1}}$ & $\frac{\Delta G_{\xi}}{{\rm kJ\,mol}^{-1}}$ & $R_{\rm min}/$\AA \\ 
1663: \hline 
1664: Na$^+$ & 1 &-398 &-334     &-394 &1.83 \\ 
1665: K$^+$ & 1 & -271 &-246     &-282 &2.35 \\ 
1666: Ca$^{2+}$ & 2 &-1306 &-1181&-1364 &2.09 \\ 
1667: F$^-$ & -1 & -580 &-274    &-630 &2.25 \\ 
1668: Cl$^-$ & -1 &-371 &-198    &-342 &2.97 \\ 
1669: Br$^-$ & -1 & -358 &-192   &-328 &3.02 \\
1670: \end{tabular}
1671: \caption{Solvation free energies for charged LJ spheres in SPC water
1672: from the simulations of Hummer {\it et al.}\cite{hummer:jpc:1996}
1673: compared to the theoretical result $\Delta G$. For $\delta$ we use the
1674: best fits $\delta_{\rm bf}$ to the solvation of neutral spheres as
1675: shown in Tab.~I. $\Delta G_{\xi}$ is the result when the dielectric
1676: boundary shift $\xi$ is applied, see text.}
1677: \label{tab1}
1678: \end{center}
1679: \end{table}
1680: 
1681: \begin{table}
1682: \begin{center}
1683: \begin{tabular}{l | c c c c c}
1684:   System & $\delta/{\rm \AA}\;\;\;$ & vdW attraction & $\;\;\;Z$ & $W(s_0)/k_BT$ & dewetted\\ 
1685: \hline 
1686: I   & 0.00 & no & 0 & -57.6 &yes \\ 
1687: II  & 0.75 & no & 0 & -34.1 &yes \\ 
1688: III & 0.75 & yes & 0 & -6.3 &yes \\ 
1689: IV  & 0.75 & yes & 4 & -9.2 &yes \\ 
1690: V   & 0.75 & yes & 5 & -5.1 &no \\ 
1691: VI  & 0.75 & yes & 1 (oc)& -1.3 &no \\
1692: \end{tabular}
1693: \caption{Studied systems for two alkane-assembled spherical
1694: solutes. $W(s_0)$ is the inter-solute pmf. If $r(z=0)\neq0$ the system
1695: is 'dewetted'. In system VI the solutes' charge is located off-center
1696: (oc) at the solute surface.}
1697: \label{tab4}
1698: \end{center}
1699: \end{table}
1700: 
1701: \clearpage 
1702: 
1703:  \begin{figure}[htb]
1704:  \begin{center}
1705: %    \epsfig{file=fig1.ps, width=7cm, angle=-90}
1706:    \caption{The particular solvation energy contributions $\Delta
1707:    G_i(R)$ with $i={\rm p,int,ne}$ in Eq.~(\ref{eq:sphere}) for one LJ
1708:    sphere with Na$^0$ parameters given in Tab.~I. The pressure term
1709:    $\Delta G_{\rm pr}$ (thin solid line) with $P=1$atm is basically
1710:    zero on this scale. The interfacial term $\Delta G_{\rm int}(R)$
1711:    (dotted line) with $\gamma_{\rm lv}=65$mJ/m$^2$ increases with
1712:    radius $R$. The LJ term $\Delta G_{\rm ne}(R)$ is given by the
1713:    dashed line. The sum of the three contribution gives the total
1714:    $\Delta G(R)$ (solid line) with a minimum at $R_{\rm min}=2.32$\AA~
1715:    for the uncharged sodium Na$^0$. The inset shows the electrostatic
1716:    contribution $\Delta G_{\rm es}(R)$ (dot-dashed line) and the total
1717:    $\Delta G(R)$ for the charged Na$^+$ with a minimum at $R_{\rm
1718:    min}=1.83$\AA. The best-fit Tolman length is $\delta_{\rm
1719:    bf}=0.79$\AA.}
1720:  \label{fig:G}
1721:  \end{center}
1722:  \end{figure}
1723: 
1724:  \begin{figure}[htb]
1725:  \begin{center}
1726: %    \epsfig{file=fig2.ps, width=7cm, angle=-90}
1727:    \caption{Mean $H(z)$ and Gaussian $K(z)$ curvature and shape
1728:    function $r(z)$ (solid lines) for ethane. The canonical SAS (dashed
1729:    line) from rolling a probe sphere with radius $r_{\rm p}=1.4$\AA~ over
1730:    the vdW surface (shaded region) is also shown. The vdW surface is
1731:    defined by the solute-solute LJ-radius $\sigma_{ss}/2=1.73$\AA.\cite{sigmanote}}
1732:  \label{fig:ethane}
1733:  \end{center}
1734:  \end{figure}
1735: 
1736:  \begin{figure}[htb]
1737:  \begin{center}
1738: %    \epsfig{file=fig3.ps, width=7cm, angle=-90}
1739:    \caption{Mean $H(z)$ and Gaussian $K(z)$ curvatures and shape
1740:    function $r(z)$ for two alkane-assembled solutes of radius
1741:    $R_0=15$\AA~ (shaded region) at a distance $s_0=8$\AA~ for systems
1742:    I-VI. The position of the charges $Z=\pm 1$ in VI are indicated by
1743:    arrows. Curvatures are not shown for the 'wet' systems V and VI. }
1744:  \label{fig:nano}
1745:  \end{center}
1746:  \end{figure}
1747: 
1748:  \begin{figure}[htb]
1749:  \begin{center}
1750: %   \epsfig{file=fig4.ps, width=7cm, angle=-90} 
1751: %   \epsfig{file=fig5.ps,width=7cm, angle=-90}
1752:    \caption{Top frame: theoretical pmfs for the systems I-III, and VI
1753:    versus the solute distance $s_0$. Bottom frame: corresponding mean
1754:    forces.}
1755:  \label{fig:pmf}
1756:  \end{center}
1757:  \end{figure}
1758: 
1759: 
1760: \begin{figure}[htb]
1761:  \begin{center}
1762: %   \epsfig{file=fig6.ps,width=7cm, angle=-90}
1763:    \caption{Mean force $\beta F$\AA~ between to nanosized solutes
1764:    versus surface-to-surface distance $s_0$. The symbols denote the MD
1765:    simulation results from Dzubiella {\it et
1766:    al.}\cite{dzubiella:jcp:2003,dzubiella:jcp:2004} for neutral
1767:    spheres with radius $R_0=12$\AA~ (circles) and $R_0=10$\AA
1768:    (squares), and oppositely charged spheres with radius $R_0=10$\AA~
1769:    and charge $Q=2$e (diamonds) and $Q=5$e (triangles). The
1770:    corresponding theoretical results using $\delta=0.9$\AA~ are shown
1771:    by solid lines; the range of the strong hydrophobic attraction
1772:    decreases with decreasing radius and increasing charge. Dotted
1773:    lines through the symbols are guides for the eye. The dashed line
1774:    is the theory for $R_0=12$\AA~ and $\delta=0.75$\AA.}
1775:  \label{fig:mf}
1776:  \end{center}
1777:  \end{figure}
1778: 
1779: 
1780:  \begin{figure}[htb]
1781:  \begin{center}
1782: %   \epsfig{file=fig7.eps, width=7cm, angle=0}
1783:    \caption{Distribution of the dielectric constant in space for two
1784:    nanosolutes with $R_0=15$\AA~ at a distance $s_0=7$\AA~ carrying a
1785:    charge $Q=5e$ (system V). The region between the spheres is
1786:    dewetted. The distribution is scaled by $\epsilon_l=78$.}
1787:  \label{eps}
1788:  \end{center}
1789:  \end{figure}
1790: 
1791: \clearpage
1792: 
1793: {\Large Figure 1}
1794:  \begin{figure}[htb]
1795:  \begin{center}
1796:     \epsfig{file=fig1.ps, width=12cm, angle=-90}
1797:  \end{center}
1798:  \end{figure}
1799: 
1800: \clearpage
1801: {\Large Figure 2}
1802:  \begin{figure}[htb]
1803:  \begin{center}
1804:     \epsfig{file=fig2.ps, width=12cm, angle=-90}
1805:  \end{center}
1806:  \end{figure}
1807: 
1808: \clearpage
1809: {\Large Figure 3}
1810:  \begin{figure}[htb]
1811:  \begin{center}
1812:     \epsfig{file=fig3.ps, width=12cm, angle=-90}
1813:  \end{center}
1814:  \end{figure}
1815: 
1816: \clearpage
1817: {\Large Figure 4}
1818:  \begin{figure}[htb]
1819:  \begin{center}
1820:     \epsfig{file=fig4a.ps, width=10cm, angle=-90} 
1821:     \epsfig{file=fig4b.ps,width=10cm, angle=-90}
1822:  \end{center}
1823:  \end{figure}
1824: 
1825: \clearpage
1826: {\Large Figure 5}
1827: \begin{figure}[htb]
1828:  \begin{center}
1829:     \epsfig{file=fig5.ps,width=12cm, angle=-90}
1830:  \end{center}
1831:  \end{figure}
1832: 
1833: \clearpage
1834: {\Large Figure 6}
1835:  \begin{figure}[htb]
1836:  \begin{center}
1837:     \epsfig{file=fig6.eps, width=12cm, angle=0}
1838:  \end{center}
1839:  \end{figure}
1840: 
1841: 
1842: \end{document}
1843: 
1844: 
1845: 
1846: 
1847: 
1848: 
1849: