cond-mat0405292/jcp.tex
1: %\documentclass[prb,preprint,groupedaddress]{revtex4}
2: %\documentclass[aps,preprint,superscriptaddress]{revtex4}
3: \documentclass[prb,twocolumn,groupedaddress]{revtex4}
4: %\documentclass[aps,twocolumn,superscriptaddress,s howpacs]{revtex4}
5: %\documentclass[twocolumn,prb,groupedaddress,showpacs,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: \usepackage{amsfonts}
9: \usepackage[dvips]{graphicx}
10: \usepackage[]{caption}
11: \usepackage[]{epsfig}
12: \bibliographystyle{apsrev}
13: \voffset 2cm 
14: \newcommand {\Tp}{{\rm T}_+}
15: \newcommand {\Tm}{{\rm T}_-}
16: \newcommand {\Tn}{{\rm T}_0}
17: 
18: \newcommand {\Cn}{{\rm C}_0}
19: \newcommand {\Cp}{{\rm C}_+}
20: \newcommand {\Cm}{{\rm C}_-}
21: 
22: \newcommand {\Sn}{{\rm S}_0}
23: \newcommand {\Sp}{{\rm S}_+}
24: \newcommand {\Sm}{{\rm S}_-}
25:  
26: 
27: \begin{document}
28: 
29: \title{Competition of hydrophobic and Coulombic interactions between nano-sized solutes}
30: 
31: \author{J. Dzubiella}
32: \email[e-mail address:] {jd319@cam.ac.uk}
33: \affiliation{University Chemical Laboratory,
34: Lensfield Road,
35: Cambridge CB2 1EW,
36: United Kingdom}
37: \author{J.-P. Hansen}
38: \affiliation{University Chemical Laboratory,
39: Lensfield Road,
40: Cambridge CB2 1EW,
41: United Kingdom}
42: \date{\today}
43: 
44: \begin{abstract}
45: The solvation of charged, nanometer-sized spherical solutes in water,
46: and the effective, solvent-induced force between two such solutes are
47: investigated by constant temperature and pressure Molecular Dynamics
48: simulations of model solutes carrying various charge patterns. The
49: results for neutral solutes agree well with earlier findings, and with
50: predictions of simple macroscopic considerations: substantial
51: hydrophobic attraction may be traced back to strong depletion
52: (``drying'') of the solvent between the solutes. This hydrophobic
53: attraction is strongly reduced when the solutes are uniformly charged,
54: and the total force becomes repulsive at sufficiently high charge;
55: there is a significant asymmetry between anionic and cationic solute
56: pairs, the latter experiencing a lesser hydrophobic attraction. The
57: situation becomes more complex when the solutes carry discrete (rather
58: than uniform) charge patterns. Due to antagonistic effects of the
59: resulting hydrophilic and hydrophobic ``patches'' on the solvent
60: molecules, water is once more significantly depleted around the
61: solutes, and the effective interaction reverts to being mainly
62: attractive, despite the direct electrostatic repulsion between
63: solutes. Examination of a highly coarse-grained configurational
64: probability density shows that the relative orientation of the two
65: solutes is very different in explicit solvent, compared to the
66: prediction of the crude implicit solvent representation.  The present
67: study strongly suggests that a realistic modeling of the charge
68: distribution on the surface of globular proteins, as well as the
69: molecular treatment of water are essential prerequisites for any
70: reliable study of protein aggregation.
71: \end{abstract}
72: 
73: %\pacs{82.70.Uv,87.16.Ac,61.20Ja,68.08.Bc}
74: 
75: \maketitle
76: \section{Introduction}
77: It is a well-known fact of physical chemistry that solvophobic solutes
78: of similar sizes and shapes tend to attract each other in an
79: incompatible solvent. Classic examples are the effective attraction
80: between monomers of polymer coils in poor solvent, which leads to
81: collapse below the $\Theta$-temperature,\cite{rubinstein} or the
82: attraction between hydrophobic surfaces in
83: water.\cite{chandler_review} The effective attraction ultimately leads
84: to phase separation of the solvent and solute as the concentration of
85: the latter increases. The solvent-averaged effective interaction (or
86: potential of mean force) is related to the variation of free energy
87: upon bringing the two solutes from infinite to a finite separation in
88: the solvent. The change in free energy has an entropic
89: component, associated with the reorganization of the solvent molecules
90: around the two solutes, and an energetic contribution
91: which accounts for the deficit in attractive interactions between
92: solvent molecules close to the solutes. There is some analogy between
93: solvophobic attraction and the well-known depletion interaction
94: between colloidal particles induced by a depletant like non-adsorbing
95: polymers.\cite{likos:physrep} In the latter case the depletion
96: attraction can be essentially understood in terms of excluded volume,
97: and is hence of entropic origin, while hydrophobic interactions have a
98: large energetic contribution, associated with the
99: formation or break up of hydrogen bonds.
100: 
101: It has been recognized that the size of the solute plays an important
102: role in understanding its solvation energy, and effective
103: solute-solute attraction.\cite{chandler_review} For solutes of a
104: characteristic size larger than a few molecular diameters (typically
105: larger than 1nm), a mechanism first envisioned by Stilinger
106: \cite{stilinger} is that of solvent dewetting (``drying''), i.e. the
107: solvent molecules tend to move away from the surface of a large
108: solute, and form a liquid-gas like interface parallel to the solute
109: interface.\cite{chandler_review, hummer:prl} The overlap of the drying
110: zones associated with two large solutes as their surfaces come
111: together may then give rise to an effective attraction,\cite{lum:jpc}
112: very much like the depletion mechanism between-colloidal
113: particles. The mechanism holds, a priori, for any solvent, provided
114: that it is in a thermodynamic state close to liquid-vapor
115: coexistence.\cite{huang:jpc} The drying mechanism and resulting
116: attraction between two plate-like solutes was confirmed by Molecular
117: Dynamics (MD) simulation of Wallquist and Berne.\cite{wallquist:jpc}
118: Similar simulations were carried out for two large spherical solutes
119: in a Lennard-Jones solvent, and attractive solvation force profiles
120: were determined.\cite{kinoshita,shinto,qin}
121: 
122: However most nano-scale biomolecular solutes, like proteins, carry
123: electric charges, which make them, at least partly, hydrophilic. The
124: main objective of the present paper is to investigate the influence of
125: solute-solute and solute-solvent electrostatic interactions on the
126: effective, solvent-induced potential of mean force between two
127: solutes. In order to make contact with earlier work on neutral
128: hard-sphere solutes, we restrict the present investigation to
129: spherical solutes of identical radii $R\lesssim 1$nm, but carrying
130: various surface charge patterns. The two main questions which will be
131: addressed are: a) how does the competition between hydrophobicity and
132: electrostatics affect the total effective force between anionic and
133: cationic solutes; b) is the total force sensitive to details of the
134: charge patterns carried by the solutes? In particular are there
135: significant differences between results obtained with continuous and
136: discrete charge patterns of the solutes?  Such differences were
137: recently highlighted by a calculation of the second virial coefficient
138: in an implicit solvent model of globular proteins, which does not, of
139: course, allow for hydrophobic attraction.\cite{elshad:epl:2002} 
140: 
141: We have attempted to answer these questions by a series of constant
142: pressure MD simulations of two spherical solutes of varying radii (up
143: to $R=1.3$nm) immersed in water modeled by the SPC/E intermolecular
144: potential,\cite{berendsen:jpc} taken under normal conditions, i.e. close to
145: liquid-vapor coexistence. The paper is structured as follows. The
146: models and simulation procedures are detailed in Sec. \ref{sec:model}. The
147: solvation of single charged solutes is examined in
148: Sec. \ref{sec:single}. The effective interaction between two solutes as
149: a function of the mutual distance is estimated from a simple
150: macroscopic theory in Sec. \ref{sec:force:theory}, where the method for
151: extracting the mean effective force from simulations is also
152: defined. The results from MD simulations for several charge patterns
153: are presented in Sec. \ref{sec:force:sim}, while concluding remarks are
154: made in Sec. \ref{sec:conclusion}.
155: 
156: Part of the present results were briefly reported elsewhere.\cite{dzubiella:jcp:2003}
157: 
158: \section{Models and methodology}
159: 
160: \label{sec:model}
161: 
162: \begin{figure}
163:   \begin{center}
164: \includegraphics[width=5.0cm,angle=0.,clip]{fig1.eps}
165:     \caption{Sketch of the SPC/E water model. The vector $\vec\omega$
166:     embodies the orientation of the molecule.}
167: \label{fig:spce} 
168: \end{center}
169: \end{figure}
170: 
171: 
172: The MD simulations were carried out on periodic samples containing
173: $N_{\rm w}$ water molecules and one or two solutes.  The SPC/E model
174: of a water molecule \cite{berendsen:jpc} is sketched in Fig.~1. Two
175: water molecules interact via a Lennard-Jones potential between the
176: oxygen (O) sites, and the bare Coulomb potentials between the 9 pairs
177: of sites. The Lennard-Jones parameters are
178: $\epsilon=0.6502$kJmol$^{-1}$ and $\sigma=3.169$\AA. The molecules are
179: assumed to be rigid (with OH bond lengths and HOH bond angle specified
180: in Fig.~1) and nonpolarizable.  The solutes are smooth spheres of
181: bare radius $R_{0}$, which interact with the O-site of the water
182: molecules by the purely repulsive potential
183: \begin{eqnarray}
184: V_{0}(r) = \phi (r-R_{0})^{-12},
185: \label{eq:solutepot}
186: \end{eqnarray}  where $r$ is the distance from the solute center to
187: the O-site, and the energy scale $\phi$ is chosen such that the O-atom
188: experiences a repulsive energy $k_{\rm B}T$ at a distance
189: $r-R_{0}=$1\AA~ from the solute surface. With this convention, the
190: effective radius of the solutes may be defined as $R=R_{0}+1$\AA.  The
191: purely repulsive interaction (\ref{eq:solutepot}) is chosen to mimic a
192: strongly hydrophobic interaction between the neutral solute and the
193: water molecules. The model involving spherical solutes with no charged
194: site will be referred to as $\Sn$.
195: 
196: Since the main objective of our work is to investigate the difference
197: in the effective, solvent induced interaction between the cases of
198: neutral and charged solutes, we have considered several models for the
199: latter (cf. Table 1). In the simplest model, a total charge $Q=qe$
200: (where $e$ is the proton charge) is assumed to be uniformly
201: distributed over the solute surface. According to Gauss' theorem, this
202: is equivalent to placing a single charged site $Q$ at the center of the
203: spherical solute. We consider both anionic ($q<0$) and cationic
204: ($q>0$) solutes and the corresponding models will be referred to as
205: $\Sm$ and $\Sp$. To ensure overall electroneutrality of the system,
206: the total charge carried by the solutes must be compensated either by
207: a uniform background of total opposite charge permeating the system,
208: or by explicitly including counterions. For the latter we choose
209: Cl$^-$ (cationic solutes) and Na$^+$ (anionic solutes) ions. Their
210: mutual interactions and coupling to water molecules involve the
211: standard Coulomb-interactions, and a Lennard-Jones part,
212: with $\epsilon$ and $\sigma$ parameters taken from
213: Spohr.\cite{spohr:1999} The short range interaction with the solutes
214: is again described by (\ref{eq:solutepot}).
215: 
216: 
217: 
218: \begin{table}
219: \begin{center}
220: %\includegraphics[width=4.0cm,angle=0.,clip]{figs/table.xfig.eps}
221: \begin{tabular}{c|c c c|c c c|c c c}
222:  & $\Sn$ & $\Sp$ & $\Sm$ & $\Tn$ & $\Tp$ & $\Tm$ & $\Cn$ & $\Cp$ & $\Cm$ \\
223: \hline
224: $N_c$           & 0   & 1  & 1  & 4 & 4 &  4 & 8  &  8 &  8 \\
225: $Q/e$           & 0   & q  & -q & 0 & 4 & -4 & 0  &  8 & -8 \\
226: $R_0/{\rm \AA}$ & $0-12$ & 10 & 10 &10 &10 & 10 & 10 & 10 & 10 \\
227: $R_c/{\rm \AA}$ & - & 0 & 0 &10 &10 & 10 & 10 & 10 & 10 \\
228: $R_{cc}/{\rm \AA}$ & - & 0 & 0 &16.33 &16.33 & 16.33 & 11.55 & 11.55 & 11.55 \\
229: \end{tabular}
230: \caption{Characteristics of the different models
231: ${\rm S}_{0\pm},{\rm T}_{0\pm},{\rm C}_{0\pm}$ used. $N_c$ is the number of charges
232: carried by the solute, while $Q$ is the net charge. $R_0$ is
233: the bare solute radius. By $R_c$ we denote the distance of the
234: charges to the center of the solute, and $R_{cc}$ quantifies the
235: nearest neighbor distance between the charges. In the charged
236: ${\rm S}$-models the charge is located in the center of the solutes, while
237: in the ${\rm T}$ and ${\rm C}$ models the charges are distributed tetrahedrally
238: and cubically on the sphere surface.  }
239: \label{tab2}
240: \end{center}
241: \end{table}
242: 
243: \begin{figure}
244:   \begin{center}
245: \includegraphics[width=7.0cm,angle=0.,clip]{fig2.eps}
246:     \caption{Solute with (a) tetrahedral (T models) and (b) cubic (C
247:     models) charge distribution. The charges are connected in this
248:     figure for a better visualization of the structure.}
249: \label{fig:proteins} 
250: \end{center}
251: \end{figure}
252: 
253: 
254: In order to investigate the sensitivity of the effective forces to
255: details of the solute charge patterns, we have also considered models
256: with discrete charge distributions involving $N_c$ point charges
257: placed on the surface of the solutes. In the tetrahedron model,
258: $N_c=4$ charges are tetrahedrally arranged at a distance $R_0$ from
259: the center of the solutes, as illustrated in
260: Fig.~\ref{fig:proteins}(a); we consider both the cases where all 4
261: charges are of the same sign ($q=\pm 4$, referred to as models $\Tp$
262: and $\Tm$), and the neutral situation, where two charges are positive
263: and two are negative (model $\Tn$). We have also considered cubic
264: charge distributions, as sketched in Fig.~\ref{fig:proteins}(b),
265: where $N_c=8$. In models $\Cp$ and $\Cm$ all 8 charges are positive
266: ($q=8$) or negative ($q=-8$), whereas in model $\Cn$, four vertices
267: carry a charge $+e$, while the other four carry opposite charges in
268: an alternating arrangement such that the three nearest neighbors of a
269: negative charge are positive and vice versa. A similar model of
270: globular proteins was considered by Allahyarov {\it et
271: al.},\cite{elshad:epl:2002} but in an implicit (continuous) solvent
272: representation. 
273: 
274: In order to avoid very close approaches of water
275: H-sites and the solute surface charges, the charged sites at the
276: vertices of the tetrahedron or cube, situated at a distance $R_0$ from
277: the solute center, are not simply point charges, but are modeled by
278: Cl$^-$ or Na$^+$ ions. The corresponding LJ potentials prevent these
279: sites and the water H atoms to come too close, and hence unreasonably
280: large electrostatic forces, which could lead to electrostatic
281: ``sticking'' of the water molecules to the solute surface.
282: The total interaction energy between one solute and the $N_{\rm w}$ water
283: molecules in a periodically repeated, cubic simulation cell is:
284: \begin{eqnarray}
285: V_{\rm sol} &  = & \sum_{i=1}^{N_{\rm w}} V_{\rm 0}(r_{i})+
286: \sum_{i=1}^{N_w}\sum_{\alpha=1}^{N_{\rm c}} V_{\rm
287:   LJ}(r_{i}^{\alpha 1})\nonumber \\
288: & + &\sum_{i=1}^{N_w}\sum_{\alpha=1}^{N_{\rm c}}\sum_{\beta =1}^3 {q_\alpha}{q_{\beta}}\phi_{\rm EW}(\vec r_{i}^{\alpha\beta}),
289: \label{eq:totalsolutepot}
290: \end{eqnarray}
291: where $r_i$ is the distance from the center of the solute to the
292: O-atom of the $i$th water molecule, and $r_{i}^{\alpha\beta}$ is the
293: distance from site $\alpha$ on the solute to site $\beta$ of the $i$th
294: water molecule ($\beta=1$ for the oxygen site). The first term on the
295: rhs of Eq.~(\ref{eq:totalsolutepot}) corresponds to the short ranged
296: repulsion (\ref{eq:solutepot}); the second term is the sum of
297: Lennard-Jones interactions between all $N_c$ sites of the solute and
298: the O-sites ($\beta=1$) of the $N_{\rm w}$ water molecules, which
299: depend on the corresponding site-site distance $r_i^{\alpha 1}$;
300: finally the last term accounts for the Coulombic interactions between
301: all $N_c$ solute sites and all 3 water sites; $\phi_{\rm EW}(\vec r)$
302: is the electrostatic interaction between two elementary charges,
303: properly summed over an infinite array of periodic images, using the
304: smooth-particle-mesh Ewald (SPME) method \cite{essmann:jcp} (see the
305: Appendix A for details). An expression similar to
306: Eq.~(\ref{eq:totalsolutepot}) holds for the total interaction between a
307: solute and its Cl$^-$ or Na$^+$ counterions.
308: 
309: The MD simulations were carried out with the DLPOLY2 \cite{dlpoly}
310: package, using the Verlet leapfrog algorithm,\cite{frenkelsmit} with
311: a timestep of 2fs. Simulations were carried out at constant pressure
312: ($P=1$atm) and constant temperature ($T=300$K), using appropriate
313: barostats and thermostats (see Appendix A). We emphasize the importance
314: of using constant pressure simulations of charged, aqueous systems:
315: electrostriction and drying mechanisms modify the density of water in
316: a finite, closed system in a significant way. In the present $NPT$
317: ensemble simulations, the overall density of water varied from
318: $\rho_0=0.033{\rm \AA}^{-3}$ to $0.035{\rm \AA}^{-3}$; in estimating
319: the average water density, the effective volume $4\pi R^3/3$ of the
320: solutes must be subtracted. The choice of the box length $L$ of the
321: periodically repeated simulation cell and the arrangement of solutes
322: inside the cell are discussed in the Appendix A. The cell contained up
323: to $N_{\rm w}=3000$ water molecules.
324: 
325: 
326: 
327: 
328: 
329: 
330: 
331: \section{Solvation of a single charged solute}
332: \label{sec:single}
333: 
334: Before embarking on the main subject of this paper, namely the
335: water-induced effective interaction between two protein-like solutes,
336: we first consider the solvation of a single, neutral or charged
337: spherical solute, a problem which has been abundantly addressed in the
338: literature, since the pioneering work of Born \cite{born} on solvation
339: free energies of ions, and of Reiss {\it et al.} on the
340: scaled-particle theory of cavity formation and hard sphere solutes.
341: \cite{reiss:jcp:1959}
342: 
343: 
344: \subsection{Water structure around a solute}
345: \label{sec:structure}
346: 
347: Consider first the structure of water around an isolated neutral
348: solute (model $\Sn$). Fig.~\ref{fig:profiles_neutral} shows MD results
349: for the oxygen and hydrogen radial distribution functions (RDF), or
350: density profiles, around a solute for various solute radii $R$. The
351: height of the main peak in the RDF of both O and H sites of the
352: water molecules is seen to first grow with increasing $R$, and this may
353: be rationalized in terms of enhanced packing of the molecules at the
354: surface of the solute. But for $R\gtrsim 5$\AA, the peak height is
355: seen to decrease monotonically, due to the unbalanced attraction
356: experienced by the water molecules near the surface from bulk
357: water. Very similar predictions of the depletion of water around large
358: spherical solutes (or cavities) have been reported earlier in the
359: literature.\cite{stilinger,huang:jpc,lum:jpc} The hydrogen and oxygen
360: peaks are located at nearly the same distance from the solute for any
361: given radius $R$, with a tendency of the hydrogen peak to be slightly
362: further out. This seems to indicate that there is no strong
363: orientation of the water molecules in the first solvation shell
364: towards or away from the solute surface.
365: 
366: \begin{figure}
367:   \begin{center}
368: \includegraphics[width=8cm,angle=0.,clip]{fig3.eps}
369:     \caption{Density profiles of water oxygen and hydrogen atoms
370:       (inset) around one isolated neutral solute ($\Sn$) plotted for
371:       different radii $R/{\rm \AA}=2,3,4,5,6,11$.}
372: \label{fig:profiles_neutral} 
373: \end{center}
374: \end{figure}
375: 
376: \begin{figure}
377:   \begin{center}
378: \includegraphics[width=8cm,angle=0.,clip]{fig4.eps}
379:     \caption{Orientation parameter $P(r)$ defined in
380:     Eq.~(\ref{eq:orient}) of the water particles around a solute with
381:     radius $R=11$\AA~ carrying no charge ($\Sn$, solid line), and
382:     charge $q=\pm 5$ (${\rm S}_\pm$, dashed lines), and $q=\pm 10$
383:     (${\rm S}_\pm$, dotted lines).}
384: \label{fig:orient} 
385: \end{center}
386: \end{figure}
387: 
388: \begin{figure}
389:   \begin{center}
390: \includegraphics[width=8cm,angle=0.,clip]{fig5.ps}
391:     \caption{Density profiles of water oxygen and hydrogen atoms
392:       (inset) around a (a) positively ($\Sp$) and (b) negatively
393:       ($\Sm$) charged solute with radius $R=11$\AA~ and central charge
394:       $q=\pm 5$ (dashed line), and $q=\pm 10$ (dotted line). We also
395:       plot the result for the $\Sn$ model (solid line) with the same
396:       radius.}
397: \label{fig:profiles_charged} 
398: \end{center}
399: \end{figure}
400: 
401: 
402: This observation may be quantified by considering the following
403: orientational order parameter:
404: \begin{eqnarray}
405: P(r)=\left<\frac{\vec\omega\cdot\vec r}{|\vec\omega||\vec r|}\right>_r,
406: \label{eq:orient}
407: \end{eqnarray}
408: where $\vec \omega$ is the water molecule orientation vector
409: (cf. Fig.~1), and the configurational average is taken for a fixed
410: distance $r$ from the solute center to the O-site of the water
411: molecules. MD results for a neutral solute of radius $R=11$\AA~ are
412: plotted in Fig.~\ref{fig:orient}.  The curve $P(r)$ takes slightly
413: positive values ($P\simeq 0.05$) for water in the first solvation
414: shell ($r\simeq 12.5$\AA), indicating a weak tendency of the hydrogen
415: atoms to point away from the solute.
416: 
417: The effect of charging the solute is illustrated in
418: Fig.~\ref{fig:profiles_charged}(a) and (b), where the RDFs are plotted
419: for a fixed radius $R=11$\AA, and charges $q=0,\pm 5$, and $\pm 10$,
420: within the $\Sn$ and ${\rm S}_\pm$ models (neutral, or uniformly
421: charged solutes). For positive charges,
422: Fig.~\ref{fig:profiles_charged}(a), the height of the first peak in
423: both oxygen and hydrogen RDFs is seen to shift to shorter distances,
424: and to increase as $q$ increases, signaling an effective attraction of
425: the dipolar solvent molecules due to the radial electric field. The
426: trend is similar for negative charges, as regards the oxygen
427: RDF. However the initial first peak (at $r\simeq 12$\AA~ for $q=0$) in
428: the hydrogen RDF is seen to split into a prepeak around $r\simeq
429: 10$\AA~ and a broad feature close to the initial peak when
430: $q=-5$. This points to a reorientation of the water molecules in the
431: first hydration shell, with the positive hydrogen atom preferring
432: being closer to the surface of the anionic solute. Further decrease of
433: the negative charge $(q=-10)$ consolidates this structure, with two
434: hydrogen peaks growing in amplitude
435: (cf. Fig.~\ref{fig:profiles_charged}(b)). Interestingly, while the
436: hydrogen RDFs are very sensitive to the sign of the solute charge
437: ($\Sp$ versus $\Sm$), the amplitudes of the first peaks of the oxygen
438: RDFs are nearly independent of this sign, but the peak around the
439: positive solute appears to be broader, signaling a larger water
440: coordination number in the first solvation shell. The corresponding
441: orientational order parameter $P(r)$ is plotted in
442: Fig.~\ref{fig:orient}. The absolute value of $P(r)$ has maxima at
443: contact $r\simeq 11$\AA~ and approximately one water diameter further
444: away ($r\simeq 13.5$\AA), and increases with absolute charge,
445: irrespective of the sign of the charge carried by the solute. Closer
446: inspection of the curves in Fig.~\ref{fig:orient} reveals, however, a
447: significant asymmetry, if not in the overall shape of $P(r)$, at least
448: in the amplitudes, which are typically $20\%$ larger for the negative
449: solute.  Anion/cation hydration asymmetry had already been reported
450: for microscopic ions in aqueous solution.\cite{latimer:jcp:1939}
451: 
452: \begin{figure}
453:   \begin{center}
454: \includegraphics[width=8cm,angle=0.,clip]{fig6.ps}
455:     \caption{Oxygen density profiles around the tetrahedral (a)
456:     positive $\Tp$ and (b) negative $\Tm$ solutes. The curves are
457:     for a full angular average (long dashed lines), cone averages
458:     around the charges (circles), and averages of water excluded by
459:     these cones (squares), as explained in section
460:     \ref{sec:structure}.  We also plot the density profile around a
461:     neutral solute $\Sn$ (solid line) for
462:     comparison. The inset in (a) sketches a two dimensional projection
463:     of a tetrahedral solute. The cone averages are performed over the
464:     water molecules in the grey cones (opening angle
465:     $\Theta=30^\circ$) containing a surface charge (dark circle).}
466: 
467: \label{fig:profiles_proteins} 
468: \end{center}
469: \end{figure}
470: 
471: \begin{figure}
472:   \begin{center}
473: \includegraphics[width=8cm,angle=0.,clip]{fig7.eps}
474:     \caption{Orientation profiles $P(r)$ around the tetrahedral
475:     positive $\Tp$ (circles) and negative $\Tm$ (no symbols)
476:     solutes. Cone averages around charges (solid lines) are compared to
477:     averages of water molecules excluded by those cones (long dashed
478:     lines).}
479: \label{fig:orient_protein} 
480: \end{center}
481: \end{figure}
482: 
483: 
484: We now turn to the structure of water around a solute with an
485: inhomogeneous charge distribution, restricting the discussion to the
486: tetrahedral $\Tp$ and $\Tm$ models. In order to characterize the
487: anisotropy of the problem, it is desirable to distinguish between
488: water molecules close to the four surface charges, and the remaining
489: water surrounding the solute. We achieve this by averaging over water
490: molecules whose centers fall either inside or outside well-defined
491: cones whose axes coincide with the radii joining the solute center and
492: the surface charges and whose vertices coincide with the solute
493: center. A sketch of the two-dimensional projection of one of the sides
494: of the tetrahedron and its associated cones is shown in the inset to
495: Fig.~\ref{fig:profiles_proteins}(a). Averages are taken over cones of
496: opening angle $\Theta=30^\circ$, high enough to accommodate the first
497: two solvation shells around a surface charge. The results for the
498: oxygen density profiles for water molecules inside or outside the
499: cones are shown in Figs.~\ref{fig:profiles_proteins}(a) and (b) for
500: the $\Tp$ and $\Tm$ models, respectively. The water molecules inside
501: the cones exhibit a typical solvation shell structure with a large
502: first peak, and a much lower second peak, which is hardly visible in
503: the $\Tp$ case. Outside the four cones, water appears to be highly
504: depleted compared to its distribution around a neutral $\Sn$ solute,
505: up to a radial distance $r\simeq 13$\AA. Thus the 25$\%$ of the solute
506: surface area inside the cones act as hydrophilic ``patches'' while the
507: remaining 75\% are hydrophobic. Interestingly, if an angular average
508: is taken over the total solute area, the mean density of water close
509: to the surface ($\lesssim 13$\AA) of a $\Tp$ or $\Tm$ solute is
510: significantly smaller than the corresponding density around a neutral
511: $\Sn$ solute ! Integration of the water density profile up to
512: $r=13$\AA~ yields coordination numbers (numbers of water molecules) of
513: 133 for $\Sn$, 92 for $\Tp$, and 95 for $\Tm$, showing a 30$\%$
514: depletion of water around the T-solutes compared to $\Sn$. The
515: orientational order parameter (3) for the $\Tp$ and $\Tm$ models are
516: plotted versus $r$ in Fig~\ref{fig:orient_protein}. The orientational
517: order of water molecules inside the cones is seen to be similar to
518: that around homogeneously charged solutes $\Sp$ or $\Sm$
519: (cf.~Fig.~\ref{fig:orient}). In the depleted volumes outside the cones
520: the water molecules shows little orientational order; if anything they
521: tend to orient in the direction opposite to the mean orientation
522: inside the cones.
523: 
524: Qualitatively similar observations hold for the distribution of water
525: molecules around solutes with cubic charge distribution (models $\Cp$
526: or $\Cm$), but obviously the volumes depleted of water are now
527: smaller, since the ``hydrophobic patches'' have shrunk now to only
528: half of the solute surface area. The MD simulations for the solutes
529: with non-vanishing net charge were carried out with explicit
530: counterions. Test runs where these counterions were replaced by a
531: uniform neutralizing background showed no differences within the
532: statistical uncertainties.
533: 
534: 
535: 
536: \subsection{Solvation free energy}
537: \label{sec:solvation}
538: 
539: The solvation free energy is equal to the reversible work required for
540: transferring a solute from vacuum into a solvent. For neutral solutes
541: in water, the solvation free energy is generally
542: positive,\cite{hummer:jpc:1996,lum:jpc} and for atomic-size solutes,
543: it stems mainly from the entropy cost of the restructuring water
544: molecules around the solute. For larger spherical solutes, a cross
545: over in the variation of the solvation free energy with radius occurs
546: typically around 1 nm.\cite{lum:jpc} The classic Born model
547: \cite{born} provides the simplest approach to the solvation free
548: energy of charged spherical solutes. The solvent is treated as a
549: dielectric continuum of permittivity $\epsilon$ and the hydration free
550: energy increases quadratically with solute charge and is proportional
551: to the inverse of the Born radius $R_{\rm B}$, according to
552: \begin{eqnarray}
553:   \Delta\mu_{\rm B}=-\frac{q^2e^2}{8\pi\epsilon_0 R_{\rm B}}(1-1/\epsilon).
554: \label{eq:born}
555: \end{eqnarray}
556: Note that the solvation free energy is a difference in chemical
557: potential of the solute as it is moved from vacuum into the
558: solvent. Hydration free energies from the Born model agree well with
559: experimental values, once the unknown parameter $R_{\rm B}$ is
560: defined. The Born radius for an ion can deviate substantially from its
561: Pauling radius (a measure of the size of an
562: ion).\cite{latimer:jcp:1939}
563: 
564: We have obtained solvation free energies for our model solutes by
565: thermodynamic integration, using the general formula
566: \begin{eqnarray}
567: \Delta\mu_{\rm sim}=\int_{\lambda_{0}}^{\lambda_{1}} d\lambda \left<\frac{\partial V_{\rm sol}}{\partial \lambda}\right>_{\lambda},
568: \label{eq:ti}
569: \end{eqnarray}
570: where the coupling parameter $\lambda$ gradually ``switches on'' the
571: interaction (2) between the solute and the solvent from an initial
572: state $\lambda=\lambda_0$ (say a neutral point solute) to a final
573: state $\lambda=\lambda_1$ corresponding to the complete
574: solute/solvent system.  The brackets $<..>_\lambda$ denote a
575: statistical average over all solute-solvent configurations for a
576: solute-solvent coupling characterized by $V_{\rm sol}(\lambda)$. The
577: index sim in $\Delta\mu_{\rm sim}$ indicates that the estimate of the
578: solvation free energy is based on MD simulations of a finite sample;
579: finite size corrections will be added as explained later.
580: 
581: \begin{figure}
582:   \begin{center}
583: \includegraphics[width=8cm,angle=0.,clip]{fig8.eps}
584:     \caption{Solvation free energy $\Delta\mu_0$ of a neutral
585:     spherical solute $\Sn$ in water. The inset shows $\Delta\mu_0$ divided
586:     by the sphere surface showing an asymptotic approach for large $R$
587:     to the liquid-vapor surface tension of water (dashed line),
588:     $\gamma\approx 72$mJ/${\rm m^2}$.}
589: \label{fig:freee_neutral} 
590: \end{center}
591: \end{figure}
592: 
593: In practice, we proceded in two steps. In a first stage, we computed
594: the solvation free energy $\Delta\mu_0$ of a neutral spherical solute
595: (model $\Sn$), as a function of its radius $R$. The second step is to
596: charge up the initially neutral solute to the final charge pattern. In
597: step one, the coupling parameter $\lambda$ in Eq.~(\ref{eq:ti}) is simply the
598: radius itself, and $\Delta\mu_0$ is consequently the work required to
599: blow up the solute against the normal force exerted by the solvent,
600: integrated over the particle surface; in this case the force is just
601: the radial derivative of the first term on the rhs of eq
602: (\ref{eq:totalsolutepot}). This is implemented, in practice, by
603: starting from $\lambda_0=R=0$, and increasing the radius by steps of
604: $\Delta\lambda=\Delta R=1$\AA, up to $\lambda_1=14$\AA~ (the largest
605: neutral solute considered in the present work). The averaged radial
606: force, as obtained from the MD simulations for various $R$, is
607: interpolated with a cubic spline and integrated to yield
608: $\Delta\mu_0$. The resulting solvation free energy is plotted in
609: Fig.~\ref{fig:freee_neutral}, and is seen to increase monotonically
610: with $R$. The solvation free energy per unit area, $\Delta\mu_0/4\pi
611: R^2$, is plotted in the inset to Fig.~\ref{fig:freee_neutral}, and is
612: seen to approach asymptotically a constant value for radii $R\gtrsim
613: 10$\AA; the latter is close to the liquid-vapor surface tension of
614: water, $\gamma=72$mJ/m$^2$. This behavior is close to that reported by
615: Lum {\it et al.}\cite{lum:jpc} for hard sphere cavities in water. The
616: ``softer'' solute-solvent pair potential (1) used in the present work
617: does not modify the solvation process significantly, compared to the
618: case of hard sphere solutes.
619: 
620: 
621: \begin{figure}
622:   \begin{center}
623: \includegraphics[width=8cm,angle=0.,clip]{fig9.eps}
624:     \caption{Excess solvation free energy $\Delta\mu_\pm$ of charging
625:     a spherical solute of radius $R=11$\AA~ in water homogeneously to
626:     a charge $q$ (solid line) and $-q$ (long dashed line). The symbols
627:     denote the solvation free energies of charging the ${\rm
628:     T}_{0\pm}$ and ${\rm C}_{0\pm}$ models in water. The symbols at
629:     $q=4$ are for the tetrahedron, at $q=8$ for the cube. The overall
630:     charge of the ${\rm T}_{0\pm}$ and ${\rm C}_{0\pm}$ solutes is
631:     zero (circles), positive (diamonds), and negative (squares).}
632: \label{fig:freee_charged} 
633: \end{center}
634: \end{figure}
635: 
636: 
637: In the second stage, to go from the neutral to the charged solute, the
638: charges of the $N_c$ sites on the solute are gradually turned on,
639: i.e. $q_\alpha(\lambda)=\lambda q_\alpha$ $(1\leq \alpha \leq N_c)$,
640: where $\lambda$ is varied from 0 to 1. The quantity to be averaged in
641: Eq.~(\ref{eq:ti}) is now the total electrostatic energy of the solute in
642: the field of the water molecules and their periodic images; the
643: statistical average is to be taken over all Boltzmann-weighted water
644: configurations when the electrostatic solute/solvent coupling is
645: multiplied by $\lambda$. Since for any $\lambda>0$ the system carries
646: a net charge, a compensating uniform background charge must be
647: included in evaluating the Coulombic part of
648: Eq.~(\ref{eq:totalsolutepot}) by Ewald summation. If the self
649: interaction energy of the solute with its own images and the
650: neutralizing background is properly included, the resulting free
651: energies are virtually independent of the size $L$ of the simulation
652: box,\cite{hummer:jpc:1996, hummer:jcp:1997} see the Appendix
653: B. Results for $\Delta\mu_{\pm}$ corresponding to a uniformly charged
654: solute with radius $R=11$\AA~ are plotted in
655: Fig.~\ref{fig:freee_charged} as a function of $q$, for anionic and
656: cationic solutes. Note that this is the excess free energy for
657: charging an initially neutral solute of $R=11$\AA, previously inserted
658: into the solvent, to a charge $q$. In order to obtain the total
659: solvation free energy, the contribution $\Delta\mu_0$ corresponding to
660: the insertion of the neutral solute in water ($\approx 0.6$MJ/mol for
661: $R=11$\AA) must be added to $\Delta\mu_\pm$.  The curves are
662: essentially quadratic on the scale shown, in agreement with Born
663: theory. Least squares fits of the data to the Born formula (4) yield
664: $R_{\rm B}^+=9.8$\AA~ and $R_{\rm B}^-=8.6$\AA~, both smaller than the
665: effective solute radius $R=11$\AA.  The solvation free energy of
666: cationic solutes is slightly positive for $0<q\lesssim 1$. Such a
667: behavior has already been reported for small cationic solutes (e.g
668: Na$^+$)\cite{hummer:jpc:1996,lynden-bell} and may be understood from
669: the competition between the free energy cost of the rearrangement of
670: water around the solute, and the electrostatic energy gain of the
671: dipolar solvent in the electric field of the solute. The latter
672: contribution appears to dominate already for small $|q|$ in the case
673: of anionic solutes, for which $\Delta\mu_-$ is always negative. Over
674: the whole range of absolute charge $|q|$, $\Delta\mu_-$ is
675: systematically lower than $\Delta\mu_+$, pointing to a preferential
676: solvation of anionic solutes, again in agreement with earlier findings
677: for other charged solute models.\cite{hummer:jpc:1996,lynden-bell,garde:2004}
678: In Sec. \ref{sec:structure} we learned that the restructuring of water
679: is stronger around an $\Sm$ solute than around its $\Sp$ counterpart,
680: so that one would expect a higher cost in entropy. Apparently the
681: closer approach of the hydrogen atoms to the negative solute decreases
682: the electrostatic contribution to the free energy more than the
683: positive entropic cost, resulting in an overall lower solvation free
684: energy.
685: 
686: Solvation free energies for solutes carrying discrete tetrahedral or
687: cubic charge distributions, corresponding to models $\Tp$, $\Tm$, and
688: $\Tn$, and $\Cp$, $\Cm$, and $\Cn$ are also shown in
689: Fig.~\ref{fig:freee_charged}. The contribution to the solvation free
690: energy from the steric LJ-part of the surface charge interaction with
691: the water molecules is insignificant and was neglected in the calculation
692: of $\Delta\mu_\pm$. As in the case of the uniformly charged
693: solutes, the negative solutes are preferentially solvated compared to
694: their positive counterparts. The solvation energies of the overall
695: neutral solutes $\Tn$ and $\Cn$ lie well above those of the charged
696: solutes (${\rm T}_\pm$ or ${\rm C}_\pm$). In particular $|\Delta\mu|$ of the
697: overall neutral solute with a cubic charge pattern $\Cn$ is roughly
698: three times smaller than the corresponding $|\Delta\mu_\pm|$ of the
699: globally charged solutes $\Cp$ and $\Cm$. This may be a consequence of
700: the considerable reorganization of water around the solutes with
701: surface charges of alternating sign, resulting in a substantial cost
702: in free energy. Finally, the solvation free energies of solutes with
703: discrete charge patterns (${\rm T}_\pm$ or ${\rm C}_\pm$) are seen to lie
704: 10-20$\%$ below the solvation free energies of their uniformly charged
705: counterparts ($S_\pm$ with $q=\pm 4$ and $\pm8$).
706: 
707:  
708: \section{Effective interaction between two solutes}
709: \label{sec:force:theory}
710: We now turn to the main objective of this paper, namely the
711: determination of the effective, solvent-mediated interaction between
712: two nanometer-sized neutral or charged solutes in water. In subsection
713: IV A and IV B we derive this interaction from simple macroscopic
714: considerations, while the MD methodology is presented in IV C.
715: \subsection{Phenomenological theory for charged plates}
716: A simple macroscopic argument, similar to Kelvin's theory of capillary
717: condensation predicts that water near liquid/vapor coexistence will
718: undergo ``drying'' when confined between two hydrophilic plates, below
719: a critical distance $D_c$ separating these plates.\cite{lum:jpc} We
720: extended the argument to the case of charged
721: plates,\cite{dzubiella:jcp:2003} showing that $D_c$ is strongly
722: reduced by the electrostatic energy associated with the surface charge
723: carried by the plates. The macroscopic argument is further refined
724: hereafter.  Consider two parallel plate-like solutes of area $A_1$,
725: separated by a distance $D$ and carrying opposite surface charges
726: $\pm\sigma$, immersed in a polar solvent of dielectric permittivity
727: $\epsilon$.  Neglecting edge effects (an approximation valid as long
728: as $D\ll A_1^{1/2}$, the electric field between the plates is
729: $E_{0}/\epsilon$ with $E_{0}=\sigma/\epsilon_{0}$. We require the
730: difference in the grand potential between the situations where the
731: liquid solvent ($l$) or its vapor ($g$) fill the volume $A_1 D$
732: between the two plates:
733: \begin{eqnarray}
734: \Omega_{\alpha}  =  -P_{\alpha} A_1D  + \frac{1}{2}\epsilon_{0}\frac{E^{2}_{0}}{\epsilon_{\alpha}}A_1D 
735:  +  2\gamma_{w\alpha}A_1 +\gamma_{l\alpha}A_2,
736: \label{eq:plates1}
737: \end{eqnarray}
738: where $P_{\alpha}$ is the pressure of phase $\alpha=l,g$ and
739: $\gamma_{w\alpha}$ the surface tension between phase $\alpha$ and the
740: plate (``wall"). $A_2$ is the area of the liquid-vapor interface
741: limited by the edges of the two opposite plates, which is created when
742: the volume between the plates is filled by vapor. $\gamma_{l\alpha}$
743: is the liquid-vapor surface tension when $\alpha=g$, and vanishes of
744: course when $\alpha=l$. The last term in Eq.~(\ref{eq:plates1}) may be
745: neglected for infinitely large plates.\cite{dzubiella:jcp:2003}
746: Consider a state close to phase coexistence at temperature $T$, and
747: let $\delta\mu=\mu-\mu_{\rm sat}$ be the positive deviation of the
748: chemical potential from its saturation value. Expanding the
749: $P_{\alpha}$ to linear order around their common value at saturation,
750: one arrives at the following expression for the difference in grand
751: potentials:
752: \begin{eqnarray}
753: \Omega_{g}-\Omega_{l}&=&(\rho_{ l}-\rho_{
754:   g})\delta\mu A_1D+\frac{\epsilon_{0}}{2}E^{2}_{0}(\frac{1}{\epsilon_{
755:   g}}-\frac{1}{\epsilon_{l}})A_1D\nonumber\\&+& 2(\gamma_{ w g}-\gamma_{w
756:   l})A_1-\gamma_{lg}A_2.
757: \end{eqnarray}
758: In water $\epsilon_{l}\equiv\epsilon\gg\epsilon_{g}\simeq 1$,
759: $\rho_{g}\ll\rho_{l}$ (except near critical conditions) and
760: $\gamma_{wl}-\gamma_{wg}=\gamma_{lg}\equiv\gamma$ for a purely
761: hydrophobic surface. Moreover $A_2=UD$, where $U$ is the
762: circumference of one plate. Hence:
763: \begin{eqnarray}
764: \Delta\Omega =\left(\rho_{l}\delta\mu+\frac{\epsilon_{0}}{2}E^{2}_{0}\right)A_1D- 2\gamma A_1 +\gamma UD.
765: \label{eq:deltaomega}
766: \end{eqnarray}
767: $\Delta\Omega=\Omega_{g}-\Omega_{l}$ is the reversible work bringing
768: the two plates from infinite separation (when the volume between them
769: is filled by liquid) to a distance $D$, at which ``drying'' has already
770: occurred. At contact, $\Delta\Omega(D=0)=-2\gamma A_1$, and
771: $\Delta\Omega$ increases linearly with $D$. The range of the purely
772: attractive potential is defined by the distance $D_c$ at which
773: $\Delta\Omega=0$; for $D>D_c$, the liquid is the preferred phase between the plates; for $D<D_c$, ``drying'' occurs.
774: From Eq.~(\ref{eq:deltaomega}) we obtain
775: \begin{eqnarray} 
776: D_{\rm c}\simeq\frac{2\gamma}{\rho_{l}\delta\mu+\frac{\epsilon_{0}}{2}E^{2}_{0}+\gamma U/A_1}.
777: \label{eq:Dc}
778: \end{eqnarray}
779: Near the liquid-vapor transition of water $\delta\mu\ll k_{B}T$, and
780: may be neglected compared to the surface tension term in the
781: denominator. Consider circular plates of radius $R$; then $U/A_1=2R$,
782: and if they are uncharged ($E_0=0$), $D_c=R$, which is in agreemeent
783: with previous calculations of the mean force between plate-like
784: solutes in water \cite{wallquist:jpc} and in a LJ-fluid.\cite{bolhuis}
785: In the case of high
786: surface charges ($\sigma\lesssim e/{\rm nm}^{2}$), $E_0$ can be as
787: large as $10^{10}$V/m, and the corresponding electrostatic term in the
788: denominator becomes comparable to the surface term for solute sizes of
789: a few nm. This leads to a strong reduction of $D_{\rm c}$ compared to
790: the case of neutral solutes. This reduction of $D_{\rm c}$ hints at a
791: considerable weakening of the hydrophobic interaction between two
792: solutes when the latter are charged. This trend will be confirmed by
793: the MD results in Sec.~V. Note however that the simple macroscopic
794: model ignores molecular details, and that its prediction does not, a
795: priori, apply to solutes carrying discrete charge patterns
796: (i.e. ``hydrophilic patches'') for which a more microscopic
797: description is required.
798: 
799: 
800: \subsection{Phenomenological theory for spherical solutes}
801: \label{sec:theory}
802: The previous model can be extended to the case of neutral or charged
803: spherical solutes with radius $R_0$ as follows (see
804: Fig.~\ref{fig:sketch}). 
805: When ``drying'' occurs, the simulation data
806: (cf. Fig.~\ref{fig:contour_neutral} (b) or (c)) suggest that a
807: cylindrically symmetric domain bounded by the two spherical solute
808: surfaces ($S_1$) and the curved liquid-vapor meniscus ($S_2$) is
809: filled with vapor. The surface $S_2$ is assumed to touch the solute
810: spheres tangentially (contact angle $\pi$) and to have a radius of
811: curvature $R'$. For a given surface-to-surface distance $s$ of the two
812: solutes, the volume $V$ of the ``dry'' domain and the areas $S_1$ and
813: $S_2$ are conveniently expressed in terms of the single parameter $x$,
814: as depicted in Fig.~\ref{fig:sketch}. If $x=0$, the vapor domain
815: shrinks to zero, i.e. the space between the solutes is filled with
816: liquid, while for $x=R_0$ the vapor occupies a cylindrical volume
817: $V=\pi R_0^2[(2R_0+s)-4R_0/3]$. For intermediate values of $x$, the
818: areas $S_1$ and $S_2$ are given by
819: \begin{eqnarray}
820:  S_1  (x)  &=&  4\pi R_0x \nonumber \\
821:  S_2  (x)  &=&  4\pi R'\left[h' \arcsin\left(\frac{x+s/2}{R'}\right)-(x+s/2)\right] \nonumber \\
822:  R' &  =  & R_0(x+s/2)/(R_0-x) \nonumber \\ h'  & = & \frac{R_0+R'}{R_0}\sqrt{2R_0x-x^2}.
823: \end{eqnarray}
824: \begin{figure}
825:   \begin{center}
826: \includegraphics[width=8cm,angle=0.,clip]{fig10.eps}
827:     \caption{Sketch of two spherical solutes of radius $R_0$ at a
828:     surface-to-surface distance $s$. The white volume $V$ between the
829:     solutes approximates the region depleted of water. $S_1$ and $S_2$ are
830:     the surrounding solute-vapor and liquid-vapor surface areas,
831:     resp. $R'$ is the radius of the curved surface $S_2$.}
832: \label{fig:sketch} 
833: \end{center}
834: \end{figure}
835: \begin{figure}
836:   \begin{center}
837: \includegraphics[width=8cm,angle=0.,clip]{fig11.eps}
838:     \caption{Free energy $\Delta\Omega$ of dried states between
839:     neutral spherical solutes with radius $R_0=10$ according to
840:     Eq.~(\ref{eq:domega}). The volume of the empty state is described
841:     by the parameter $x$ for a given geometry, see
842:     Fig.~\ref{fig:sketch}. The most bottom curve is for a
843:     surface-to-surface distance $s=0$, then $s$ is incremented by
844:     0.5\AA~ steps.}
845: \label{fig:omegax} 
846: \end{center}
847: \end{figure}
848: The difference in grand potential between the ``dry'' and filled
849: states is then given (in absence of the electric charges) by the
850: following generalization of Eq.~(\ref{eq:deltaomega}):
851: \begin{eqnarray}
852: \Delta\Omega=\rho_l \delta\mu V(x)-\gamma_{\rm s} S_1(x)+\gamma S_2(x),
853: \label{eq:domega}
854: \end{eqnarray}
855: where $\gamma_s$ is the surface tension of the solute-liquid interface
856: and $\gamma$ the liquid-vapor surface tension.  The first term in
857: Eq.~(\ref{eq:domega}) is the bulk free energy for creating a cavity of
858: volume $V(x)$ in water, and favors the filled state. Again, near
859: liquid-vapor equilibrium $\delta\mu\ll k_{B}T$, and the volume term
860: may be safely neglected. The second and third terms in
861: Eq.~(\ref{eq:domega}) are the surface free energies for decreasing the
862: solute-liquid and increasing the liquid-vapor interface,
863: respectively. For simplicity we first assume
864: $\gamma=\gamma_s$\cite{huang:jpc}; effects arising from
865: $\gamma\neq\gamma_s$ will be discussed later. In the following we use
866: the liquid-vapor surface tension of water $\gamma=0.174
867: k_BT{\rm\AA}^{-2}$. $\Delta\Omega(x)$ is plotted versus the geometric
868: control parameter $x$ in Fig.~\ref{fig:omegax} for a solute of radius
869: $R_0=1$nm, and various surface-to-surface distances $s$. At contact
870: ($s=0$), $\Delta\Omega(x)$ exhibits a single negative minimum at the
871: non-zero value $x=x_{\rm min}(s=0)$ signaling that the ``dry'' state
872: with volume $V(x_{\rm min})$ is stable. As $s$ increases, the global
873: minimum is raised to less negative energy values and occurs at smaller
874: values of $x$, while a second local minimum appears at $x=0$,
875: corresponding to a metastable, filled state. At the critical value
876: $s_c\simeq 3.2$\AA, the global minimum jumps from the non zero value
877: of $x$ to $x=x_{\rm min}(s_c)=0$. For $s>s_c$, the filled state is
878: favored. The effective interaction $w(s)$ between two solutes is equal
879: to the free energy difference in bringing them from infinite
880: separation to a surface-to-surface distance $s$, and is given by
881: Eq.~(\ref{eq:domega}), i.e. $w(s) =\Delta\Omega(s;x_{\rm
882: min}(s))$. The energy at the global minimum is negative for all
883: $s<s_c$ ($x_{\rm min}>0$), so that the interaction is always
884: attractive. $s_c$ is the range of the interaction; at $s=s_c$, $w
885: (s)=\Delta\Omega=0$, and $\gamma_{\rm s} S_1=\gamma S_2$. 
886: 
887: The effective potentials and forces between two solutes of different
888: radii $R_0=5,8,10,12$\AA~ from Eq.~(\ref{eq:domega}) are plotted
889: in Fig.~\ref{fig:forces_theory}. A detailed numerical investigation
890: shows that, assuming $\gamma_s=\gamma$, the range of the potential
891: (and of the resulting effective force) is $s_c\simeq 0.32R_0$, scaling
892: linearly with solute radius, independently of $\gamma$. The contact
893: value of the potential is $w(0) \simeq -1.19\gamma R_0^2$, scaling
894: with the solute surface area.  The contact value of the force is $
895: F(0)\simeq -4.28\gamma R_0$. The force increases roughly linearly with
896: a slope independent of $R_0$ (cf. Fig.~12). The results are accurately
897: represented by
898: \begin{eqnarray}
899:  F(s) =\gamma
900: \begin{cases}
901: -a_1 R_0 + a_2s & {\text {for}} \,\,s \leq s_c=a_3R_0 ; \cr
902:    0 & {\rm otherwise}. \cr
903: \end{cases}
904: \label{eq:force}
905: \end{eqnarray}
906: and
907: \begin{eqnarray}
908:  w(s) =\gamma
909: \begin{cases}
910: -a_0 R_0^2 +a_1R_0s-\frac{a_2}{2}s^2 & {\text {for}} \,\,s \leq s_c ; \cr
911:    0 & {\rm otherwise}. \cr
912: \end{cases}
913: \label{eq:pot}
914: \end{eqnarray}
915: with $a_0=1.19$, $a_1=4.28$, $a_3=0.32$, and $s_c=a_3 R_0$.
916: If the force is assumed to be linear in $s$, $a_2$ is fixed by
917: the other constants according to $a_2=2(a_1a_3-a_0)/a_3^2\simeq 3.51$.
918: The linear scaling of the contact value of the force with solute
919: radius $R$ may be rationalized by a simple consideration of the
920: potential of mean force for plates ($R=\infty$) at contact,
921: $w(0)=-2\gamma$ and an application of the Derjaguin approximation,
922: valid for weakly curved substrates (i.e. large $R$); \cite{louis:jcp}
923: this leads to the estimate $F(s=0)=-2\pi R\gamma$, which indeed
924: predicts linear scaling.
925: 
926: \begin{figure}
927:   \begin{center}
928: \includegraphics[width=8cm,angle=0.,clip]{fig12.eps}
929:     \caption{Results from the phenomenological theory in section \ref{sec:theory}
930:       for the mean force, $F^{*}=\beta F$\AA, between two neutral
931:       ($q=0$) spheres in SPC/E water. Results are plotted for solute
932:       radii $R_0=5$\AA~, $R_0=8$\AA~, $R_0=10$\AA~, and
933:       $R_0=12$\AA~. The inset shows the integrated force (potential of
934:       mean force). The depth and range of the force and potential
935:       increases with $R$.}
936: \label{fig:forces_theory} 
937: \end{center}
938: \end{figure}
939: 
940: Near a ``realistic'' solute, water will have a surface tension
941: $\gamma_s\neq\gamma$, due to Van-der-Waals and electrostatic
942: interactions, as well as the influence of curvature for small
943: solutes. One may expect $\gamma_s < \gamma$, and lowering $\gamma_s$
944: will lead to less hydrophobic attraction.  A possible
945: approach to include the electrostatic field effects due to a net
946: charge on the solutes is to absorb these effects into the
947: solute-liquid surface tension.  A naive procedure is to approximate
948: $\gamma_s$ by the solvation free energy per unit area of a charged
949: solute. In section \ref{sec:solvation} it was shown that the Born
950: expression (4) fits the MD data for uniformly charged solutes quite
951: well, once the Born radius $R_B$ has been adjusted.  Thus one may
952: write for large solutes $R\gtrsim1$nm (neglecting $1/\epsilon$ compared to 1):
953: \begin{eqnarray}
954: \gamma_s=\gamma-\frac{q^2e^2}{32\pi^2\epsilon_0 R_B^3}, 
955: \label{eq:gamma_charged}
956: \end{eqnarray}
957: which indeed lowers the hydrophobic attraction when charge is added to
958: the solutes. No hydrophobic attraction occurs when $\gamma_s=0$, so
959: that the corresponding critical charge $q_c$ satisfies:
960: \begin{eqnarray}
961: {q_c^2}={32\pi^2 e^2\epsilon_0 \gamma R_B^3}
962: \label{eq:gamma_charged}
963: \end{eqnarray}
964: For $R_B \approx 1$nm, one obtains $|q_c|\approx 3$, which in view of the
965: sensitivity to $R_B$, yields at least  the right order of magnitude, since 
966: the MD data discussed later suggest $|q_c|\approx 8$. 
967: 
968: 
969: \subsection{Forces and potentials from simulation}
970: In the MD simulations, the mean force between two solutes was
971: calculated by placing them at fixed positions ${{\vec R}_1}$ and
972: ${{\vec R}_2}$ along the body diagonal of the simulation cell, and
973: averaging over water configurations generated during the runs which
974: extended typically over 1-3 ns. The averaging was performed as long
975: the statistical error was larger than $\Delta \beta F{\rm \AA}=0.5$,
976: approximately twice the symbol size in the figures showing the
977: forces. Separate MD simulations have to be carried out for each
978: center-to-center distance ${{\vec R}_{12}}={{\vec R}_1}-{{\vec R}_2}$,
979: i.e. for a series of surface-to-surface distances $s=R_{12}-2R_0$.
980: The force acting on solute 1 is estimated from the statistical average
981: of the gradient of the total interaction energy $V_{\rm sol}$ in
982: Eq.~(\ref{eq:totalsolutepot}):
983: \begin{equation}
984: {\vec F}_1(R_{12}) = \left\langle -\nabla V_{\rm sol}(R_{12})\right\rangle_{{\vec R}_{12}},
985: \label{f1.eq}
986: \end{equation}
987: where the constrained statistical average is taken over solvent
988: configurations, when the two solutes are held fixed at a separation
989: $\vec R_{12}$. Note that while the solute translational degrees are
990: frozen, they rotate freely under the action of the torques exerted by
991: the solvent and the other solute. In other words, the calculated
992: effective forces are orientationally averaged. By symmetry, $\vec
993: F_2(\vec R_{12})=-\vec F_1(\vec R_{12})$. The magnitude of the
994: effective force is obtained by projecting onto the vector $\vec
995: R_{12}$:
996: \begin{equation}
997: F(R_{12}) = \frac{{\vec R}_1 - {\vec R}_2}{R_{12}}\cdot
998:                       {\vec F}_1(R_{12}).
999: \label{fdep.eq}
1000: \end{equation}
1001: The resulting effective solute-solute potential follows from:
1002: \begin{equation}
1003: w(R_{12})=\int_{R_{12}}^\infty F(R)dR.
1004: \label{eq:int}
1005: \end{equation}
1006: In simulations where counterions are present, the sum of all pair
1007: interactions between the latter and the solute must be added to
1008: $V_{\rm sol}$ in Eq.~(\ref{f1.eq}), i.e. the mean force acting on the
1009: solute is the statistical average of the sum of the instantaneous
1010: forces exerted by all water molecules and ions on the solute, in the
1011: presence of a fixed second solute. 
1012: 
1013: 
1014: 
1015: 
1016: \section{Molecular dynamics results for profiles and forces}
1017: \label{sec:force:sim}
1018: \subsection{Neutral solutes}
1019: 
1020: We first consider the case of uncharged solutes. The water density
1021: profiles are illustrated in Figs.~\ref{fig:contour_neutral}(a)-(e) for
1022: the case of solutes of radius $R=11$\AA~ and different
1023: surface-to-surface distances along the $z$-axis joining the
1024: centers. We plot density contours coded by variable shades of
1025: gray. The profiles are calculated using a cylindrical average around the
1026: symmetry axis (center-to-center line).  The profiles show a
1027: considerable depletion (dark region) of the solvent between the
1028: spheres, reminiscent of the observations of Wallquist and Berne for
1029: flatter solutes. \cite{wallquist:jpc} As the surface-to-surface
1030: distance $s$ is increased for fixed radius $R$, the water molecules
1031: penetrate into the region between opposite solute surfaces, as
1032: signaled by a decreasing radius of the dark  region between
1033: the spheres. Eventually at a distance between $s\approx5$\AA~ and
1034: 6\AA~ (Figs.~\ref{fig:contour_neutral}(c) and (d)) the region between
1035: the solutes fills with liquid.  When $s\gtrsim 7$\AA, the solvent
1036: layers around an isolated solute are hardly disturbed by the presence
1037: of the other solute.
1038: 
1039: Examples for the mean force for several radii 6\AA~$\leq R \leq
1040: 13$\AA~ are shown in Fig.~\ref{fig:forces_sim}. The largest radii are
1041: of the order of the size of small globular proteins or of oil-in-water
1042: micelles.  The average force obviously goes to zero at large distances
1043: $s$ and for symmetry reasons, it is directed along the
1044: center-to-center axis. As expected from a depletion mechanism, the
1045: force is attractive and its contact values and range increase with
1046: $R$. We observe for all radii that the range of the force is similar
1047: to the distance at which the drying between the solutes vanishes.  The
1048: potentials of mean force $w(s)$ may be calculated for each $R$
1049: according to Eq.~(\ref{eq:int}). The resulting potentials are shown in the inset
1050: to Fig.~\ref{fig:forces_sim}. They closely resemble results obtained
1051: for polymer-induced depletion potentials between spherical colloids,
1052: albeit on different length and energy scales.\cite{louis:jcp} The
1053: force at contact, $F(s=0)$, and the range of the force $s_c$ scales
1054: roughly with $R$ in agreement with the macroscopic model prediction
1055: (14). From the MD data we extract a rough scaling $F(s=0)\approx
1056: 3.9\gamma R$ and range $s_c\approx 0.4 R$, reasonably close to the
1057: theoretical estimates from Sec.~IV.B. Overall there is a striking
1058: qualitative and even semi-quantitative agreement between the MD forces
1059: in Fig.~14 and the predictions of tbe macroscopic model in Fig. 12.
1060: 
1061: \begin{figure}
1062:   \begin{center}
1063: \includegraphics[width=6cm,angle=0.,clip]{fig13.eps}
1064: \caption{Contour density profiles of water around two neutral
1065: $\Sn$ solutes with radius $R=11$\AA~ for surface-to-surface
1066: distances (a) $s=2$\AA, (b) $s=4$\AA, (c) $s=5$\AA, (d) $s=6$\AA,
1067: and (e) $s=7$\AA. The dark and light areas correspond to low and high
1068: densities of water molecules.}
1069: \label{fig:contour_neutral}
1070: \end{center}
1071: \end{figure}
1072: 
1073: \begin{figure}
1074:   \begin{center}
1075: \includegraphics[width=8cm,angle=0.,clip]{fig14.eps}
1076:     \caption{Simulation results (symbols) of the mean force,
1077:       $F^{*}=\beta F$\AA, between
1078:       two neutral $\Sn$ solutes in SPC/E water. The
1079:       lines are guides to the eye. Results are plotted for solute
1080:       radii  $R=6$\AA~ (squares), $R=9$\AA~
1081:       (diamonds), $R=11$\AA~ (triangles pointing up), and $R=13$\AA~ (triangles
1082:       pointing left). The inset shows the integrated force (potential of mean
1083:       force) in obvious order.}  
1084: \label{fig:forces_sim} 
1085: \end{center}
1086: \end{figure}
1087: 
1088: \subsection{Uniformly charged solutes}
1089: 
1090: 
1091: \begin{figure}
1092:   \begin{center}
1093: \includegraphics[width=8.6cm]{fig15.ps}
1094:     \caption{Density profiles of the water molecules around (a) two
1095:       neutral $\Sn$ solutes of radius $R=11$\AA~  and two oppositely
1096:       charged ${\rm S}_\pm$ solutes of radius $R=11$\AA~ carrying a
1097:       charge $\pm qe$ with (b) $|q|=2$, (c) $|q|=5$, and (d)
1098:       $|q|=10$. The surface-to-surface distance in all cases is
1099:       $s=4$\AA. In the contour plots  dark regions indicate low density
1100:       regions while high densities are plotted bright. The panels
1101:       below the contour plots show the water density $\rho$ scaled
1102:       with water bulk density $\rho_{0}$ in a cylinder of radius
1103:       $5$\AA, coaxial with the center-to-center line of the
1104:       solutes.}    
1105: \label{fig:contour_charged}
1106: \end{center}
1107: \end{figure}
1108: 
1109: \begin{figure}
1110:   \begin{center}
1111: \includegraphics[width=8cm,angle=0.,clip]{fig16.eps}
1112:     \caption{Mean force, $F^{*}=\beta F$\AA, as a function of
1113:       surface-to-surface distance $s$ for neutral $\Sn$ solutes
1114:       (circles) and oppositely charged ${\rm S}_\pm$ solutes of
1115:       $R=11$\AA~ with different central charges $\pm qe$:
1116:       $q=2$ (squares), $q=5$ (diamonds), $q=10$ (triangles up). The
1117:       dashed line represents the electrostatic force between 2
1118:       periodically repeated solutes with opposite charges $q=\pm$ 10
1119:       in a continuous solvent with permittivity $\epsilon=80$.  The
1120:       inset shows the resulting potentials of mean force; the contact
1121:       values $w(s=0)$ increase with $q$.}
1122: \label{fig:forces_opp_charged}
1123: \end{center}
1124: \end{figure}
1125: 
1126: 
1127: \begin{figure}
1128:   \begin{center}
1129: \includegraphics[width=8cm,angle=0.,clip]{fig17.eps}
1130:     \caption{Mean force, $F^{*}=\beta F$\AA, as a function of
1131:       surface-to-surface distance $s$ for equally charged ${\rm
1132:       S}_\pm$ solutes of radius $R=11$\AA~ and $q=-4$ (diamonds),
1133:       $q=4$ (circles), $q=-8$ (squares), and $q=8$ (triangles). The
1134:       inset shows the according potential of mean force.}
1135: \label{fig:forces_like_charged}
1136: \end{center}
1137: \end{figure}
1138: 
1139: 
1140: We now turn to charged solutes. Consider first oppositely charged
1141: solutes. The water density profiles are illustrated in
1142: Figs.~\ref{fig:contour_charged}(a)-(d) for the case of solutes of
1143: radius $R=11$\AA~, a surface-to-surface distance $s=4$\AA~, and
1144: various charges $q$. The upper part of each frame shows a density
1145: contour plot coded by variable shades of gray. The lower part shows
1146: density profiles along the center-to-center axis $z$, averaged over a
1147: coaxial cylindrical volume of radius $5$\AA. In Frame (a) we plot the
1148: density distribution for the neutral case $Q=0$ for comparison with
1149: the charged systems.  Frames (b)-(d) in Fig.~\ref{fig:contour_charged}
1150: show water density profiles in the vicinity of two spheres carrying
1151: opposite electric charges $\pm qe$ at their center (opposite charges
1152: ensure overall charge neutrality without any need for counterions). As
1153: $q$ increases from zero (frame (a)), water is seen to penetrate
1154: between the two solutes, the central peak around $z=0$ in the density
1155: profiles increases rapidly and its amplitude reaches roughly the bulk
1156: density of water when $q=10$. Note that this central peak is
1157: asymmetrically split, indicating the presence of two hydration layers
1158: which differ somewhat depending on their association with the anionic
1159: or cationic solute. This difference is also evident in the contact
1160: values of the outside surfaces of the solutes, and is a consequence of
1161: the different arrangements of the water dipoles around the solutes
1162: induced by the local electric fields. The asymmetry of the profiles
1163: can be rationalized by inspecting the water structure around isolated
1164: solutes, as shown in Fig.~\ref{fig:profiles_charged}(a) and (b) and
1165: Fig.~\ref{fig:orient} in Sec. \ref{sec:structure}. The hydration shell
1166: is more sharply defined around the cationic than around the anionic
1167: solute. The water dipoles tend to point radially away from the cation,
1168: while the opposite configuration is more favorable around anions.
1169: 
1170: The resulting mean forces between solutes are plotted for $q=0,2,5$
1171: and 10, as functions of the surface-to-surface distance $s$ in
1172: Fig.~\ref{fig:forces_opp_charged} together with corresponding
1173: potentials of mean force. The mean force includes the direct Coulomb
1174: interaction between the two solutes (with proper account for the
1175: periodic images), which is in fact an order of magnitude larger than
1176: the total mean force. At large distances hydrophobic interactions
1177: become negligible and the force should tend to
1178: $-q^{2}e^{2}/(4\pi\epsilon_{0}\epsilon r^{2})$, where $r=2R+s$ and
1179: $\epsilon$ is the dielectric permittivity of bulk water; the
1180: corresponding curve for $q=\pm 10$ is also shown in
1181: Fig.~\ref{fig:forces_opp_charged}.
1182: 
1183: The most striking result illustrated in
1184: Fig.~\ref{fig:forces_opp_charged} is the near independence of the
1185: force at contact, $s=0$, with respect to solute charge. From the
1186: density profiles in Fig.~\ref{fig:contour_charged} the hydrophobic
1187: attraction is expected to be reduced but this reduction is almost
1188: exactly compensated by the Coulomb attraction between solutes in the
1189: presence of the solvent. As $q$ increases, the initial slope of the
1190: effective force increases. The potential of mean force (shown in the
1191: inset to Fig.~\ref{fig:forces_opp_charged}) exhibits a contact value
1192: which increases with $q$, indicating that the reduction of hydrophobic
1193: attractive free energy clearly outweighs the increase in bare Coulomb attraction between the
1194: latter. Simulations calculating the forces at and near contact for
1195: $q=7$ and $q=15$, not shown in Fig.~\ref{fig:forces_opp_charged},
1196: confirm this trend. Note that the potential of mean force  for
1197: $q=10$ shows more long range attraction compared to the smaller $q$
1198: data due to the increased electrostatic attraction. The
1199: eye-catching kink in the force for $q=$10, at a distance $s\approx
1200: 1$\AA~ is reproducible, and is probably related to the pronounced
1201: shell structure of water molecules around highly charged solutes,
1202: discussed in sec~\ref{sec:structure}. While for neutral (and weakly
1203: charged) solutes, the O and H density profiles show little structure,
1204: they are sharply peaked at a distance $s\approx 1$\AA~ of the O atoms
1205: from the solute surface. This would lead to a complete shared
1206: hydration layer, and consequently to a kink in the force versus
1207: distance curve, between 2 flat solutes separated by $s=2$\AA~. This
1208: critical separation is shifted to shorter distances due to the
1209: curvature of spherical solutes.
1210:  
1211: In view of this delicate balance between various interactions, we have
1212: also examined the case of equally charged solutes. In this case
1213: monovalent counterions (Na$^{+}$ or Cl$^{-}$) were included to ensure
1214: overall charge neutrality. The situation is summarized in
1215: Fig.~\ref{fig:forces_like_charged} for solutes of radius $R=11$\AA~
1216: and charge $q=\pm4$ and $q=\pm8$. Charges of 4 and 8 were chosen to
1217: allow a direct comparison with the results for the models ${\rm
1218: T}_\pm$ and ${\rm C}_\pm$ in the next section.  The water density
1219: profiles are symmetric with respect to $z=0$ for equally charged
1220: solutes, but differ substantially when going from a pair of anions to
1221: a pair of cations, as discussed in Sec.~\ref{sec:structure}. This
1222: difference is reflected in the effective forces and potentials shown
1223: in Fig.~\ref{fig:forces_like_charged}. The interaction between the
1224: anionic solutes is always more attractive. For $q=\pm4$ hydrophobic
1225: attraction overcomes the repulsion between like charges, while for
1226: $q\pm8$ the electrostatic contribution dominates and the force is
1227: mainly repulsive, apart from a small attractive kink at
1228: $s=2-3$\AA. The contact value of the repulsive forces is an order of
1229: magnitude higher than in a continuous solvent, originating obviously
1230: from the lack of water between the solutes and, hence a reduced
1231: dielectric screening. 
1232: 
1233: The reduction of the hydrophobic attraction between initially
1234: uncharged solutes, upon increasing the solute charge, may be qualitatively
1235: understood by the orienting action of the strong electric field
1236: between charged solutes on the water molecules, which disrupts the
1237: local hydrogen-bond structure and moves water locally away from
1238: conditions of liquid-vapor coexistence, so that ``drying'' no longer
1239: occurs.
1240: 
1241: 
1242: \subsection{Discrete charge patterns}
1243: \label{sec:inhom}
1244: The water density distribution around two tetrahedral solutes $\Tn$
1245: (i.e carrying no net charge) is plotted in Fig.~18(a)-(e) for
1246: increasing values of the surface-to-surface distance $s$.  As in the
1247: case of neutral solutes $\Sn$, water depletion between the solutes is
1248: observable up to a solute distance of $s\simeq5$\AA. The bright
1249: regions near the surfaces indicate high density water and stem from
1250: the solvation shells in the immediate vicinity of the surface
1251: charges. We note again that the density profiles were calculated by
1252: averaging the water density cylindrically around the symmetry axis and
1253: the solutes were free to rotate in the simulations. The difference in
1254: brightness at the solute surfaces shows that certain orientational
1255: configurations are favored over others. If the tetrahedra were
1256: rotating freely (without any mutual interactions), the brightness
1257: would be the same everywhere on the solute contour, as in the case of
1258: homogeneously charged solutes. It seems that the systems chooses those
1259: configurations where the hydrophobic parts of the solutes (i.e. areas
1260: between the hydrophilic surface charges) face each other. We have also
1261: performed simulations of the overall neutral tetrahedral solute
1262: replacing the SPC/E water by a continuous solvent with permittivity
1263: $\epsilon=80$. In the latter case and for short distances, $s\lesssim
1264: 5$\AA, configurations are favored in which opposite surface charges
1265: associated with the two solutes face each other, thus strongly
1266: lowering the electrostatic energy of the system. With explicit water
1267: this is apparently no longer the case, despite the expected reduction
1268: in dielectric screening close to the surfaces; the system tries to
1269: deplete water from the solute surfaces between the solutes. For
1270: distances larger than $s\simeq5$\AA~, where no visible drying
1271: occurs, the positions of the bright regions are more smeared out on
1272: average, indicating a higher rotational freedom of the solutes. A
1273: simple configurational order parameter will be defined and discussed
1274: later.  The water profiles for the tetrahedral solutes $\Tp$ show the
1275: same behavior, in particular a visible drying for $s\lesssim5$\AA. We
1276: have not performed simulations of a pair of the negative $\Tm$
1277: tetrahedra but expect similar behavior.
1278: 
1279: \begin{figure}
1280:   \begin{center}
1281: \includegraphics[width=6cm,angle=0.,clip]{fig18.eps}
1282:     \caption{Contour density profiles of water around two $\Tn$
1283: solutes for surface-to-surface distances (a) $s=0$\AA, (b) $s=2$\AA,
1284: (c) $s=4$\AA, (d) $s=6$\AA, and (e) $s=9$\AA.}
1285: \label{fig:contour_tetra}
1286: \end{center}
1287: \end{figure}
1288: 
1289: The water density profiles around a solute $\Cn$ carrying an overall
1290: neutral cubic charge distribution are plotted in
1291: Fig.~\ref{fig:contour_cube}(a)-(c).  No water depletion is visible at
1292: any distance. In (a) the black region between the solutes comes from
1293: the fact that the solute surfaces touch.  The positions of the high
1294: density regions of water corresponding to the first solvation shells
1295: of the surface charges are on average distributed homogeneously over
1296: the sphere surface, pointing to a high orientational freedom of the
1297: solutes. The density of water close to the solute surface (bright ring
1298: in Figs.~\ref{fig:contour_cube}(a)-(c)) is on average higher compared
1299: to the tetrahedra due to the larger surface charged density. The water
1300: density profiles for the positive cubic solute $\Cp$ show a different
1301: behavior, resembling the results for the tetrahedral solute, as shown
1302: in Fig.~\ref{fig:contour_pos_cube}.  For distances $s\lesssim 4$\AA~
1303: no bright region is found between the solutes, and hence water
1304: depletion is observed. Similar to the tetrahedral case the solutes
1305: stay mainly in orientational configurations in which the hydrophobic
1306: patches face each other.
1307: 
1308: \begin{figure}
1309:   \begin{center}
1310: %\includegraphics[width=7cm,angle=0.,clip]{fig19.eps}
1311:     \caption{Contour density profiles of water around two $\Cn$
1312: solutes with cubic charge distribution for surface-to-surface
1313: distances (a) $s=0$\AA~, (b) $s=2$\AA~, and (c) $s=4$\AA~.}
1314: \label{fig:contour_cube}
1315: \end{center}
1316: \end{figure}
1317: 
1318: \begin{figure}
1319:   \begin{center}
1320: \includegraphics[width=7cm,angle=0.,clip]{fig20.eps}
1321:     \caption{Contour density profiles of water around two $\Cp$
1322: solutes for surface-to-surface distances (a) $s=0$\AA~, (b) $s=2$\AA~,
1323: and (c) $s=4$\AA~.}
1324: \label{fig:contour_pos_cube}
1325: \end{center}
1326: \end{figure}
1327: 
1328: 
1329: The effective force between the model solutes is plotted in
1330: Fig.~\ref{fig:forces_solutes}. We show the force between pairs of
1331: neutral and overall positive tetrahedra, $\Tn$ and $\Tp$ as well as
1332: between pairs of neutral and overall positive cubes, $\Cn$ and $\Cp$.
1333: Simulations with overall charged solutes were carried out with
1334: explicit counterions.  We have not performed simulations of a pair of
1335: overall negative tetrahedra and cubes. The force between pairs of
1336: overall neutral and charged tetrahedra is only slightly less
1337: attractive than between pairs of spherically symmetric neutral solutes
1338: (compare to Fig.~\ref{fig:forces_sim}). This is surprising as one
1339: expects a stronger influence of the high electric fields generated by
1340: the surface charges. We learned from the investigation of
1341: homogeneously charged solutes that an electric field can considerably
1342: lower the hydrophobic attraction. Apparently, a more anisotropic
1343: electric field distribution again favors hydrophobic
1344: attraction. Compare the force between pairs of positively charged
1345: tetrahedra and pairs of homogeneously charged solutes with the same
1346: overall charge $q=+4$ (Fig.~\ref{fig:forces_like_charged}): the
1347: attraction between the homogeneously charged solutes is less than half
1348: that between $\Tp$ solutes. As already discussed above, the strong
1349: attraction between two tetrahedra is accompanied by water depletion
1350: between them, as in the case of homogeneous neutral solutes.  The
1351: water density profile around an isolated tetrahedral solute ${\rm T}_\pm$,
1352: shown in Sec. \ref{sec:structure}, already illustrated strong
1353: depletion of water from the regions between the first solvation shells
1354: of the discrete surface charges. A possible explanation of the strong
1355: attraction between two tetrahedra is that this depletion is amplified
1356: when two hydrophobic patches of the solutes come close and face each
1357: other, and thus lowering the free energy.
1358: 
1359: 
1360: The effective force between two solutes with overall neutral cubic
1361: charge distribution shows qualitative differences compared to the
1362: tetrahedra. Cubes with zero overall charge still attract each other,
1363: but the interaction range is decreased.  Analysis of the
1364: configurations shows that for close cubic solutes ($s\approx 1-3$\AA)
1365: one positive and one negative charge belonging to different solutes
1366: are on average very close, interacting with reduced dielectric screening
1367: than in bulk water due to their mutual proximity. The attraction
1368: observed is therefore mainly due to the electrostatic contribution and
1369: not hydrophobic attraction as in the case of neutral tetrahedra. For
1370: the $\Cp$ solute the situation is again different. All charges repel,
1371: allowing the hydrophobic patches to face each other and water
1372: depletion is induced, as seen in the water profiles of
1373: Fig.~\ref{fig:contour_pos_cube}. Although equally charged, the cubic
1374: solutes still attract each other in striking contrast to the
1375: homogeneously charged solutes with $Q=8$ in explicit water
1376: (Fig.~\ref{fig:forces_like_charged}) and different to the case where
1377: water is replaced by a continuous solvent with $\epsilon=80$, also
1378: plotted in Fig.~\ref{fig:forces_solutes}.
1379:   
1380: \begin{figure}
1381:   \begin{center}
1382: \includegraphics[width=8cm,angle=0.,clip]{fig21.eps}
1383:     \caption{Mean force between  $\Tp$ (diamonds), $\Tn$ (triangles),
1384:     $\Cp$ (squares), and $\Cn$ (circles) solutes. Also plotted
1385:     is the force between two periodically repeated $\Cp$ solutes 
1386:     in a continuous solvent with $\epsilon=80$ (solid line without symbols).}
1387: \label{fig:forces_solutes}
1388: \end{center}
1389: \end{figure}
1390: 
1391: 
1392: 
1393: \begin{figure}
1394:   \begin{center}
1395: \includegraphics[width=8cm,angle=0.,clip]{fig22.eps}
1396:     \caption{Sketch of two T solutes with radius $R_0$ and a 
1397:     surface-to-surface distance $s$. The dashed lines through the
1398:     solute centers delimit a slab of width $2R_0+s$. The numbers of
1399:     solute charges (small spheres) of each solute $N_1$ and $N_2$
1400:     inside the slab define the order parameter $N=N_1 \cdot N_2$,
1401:     described in section \ref{sec:inhom}.  $N=6=3\cdot 2$ in the
1402:     configuration shown. }
1403: \label{fig:slab}
1404: \end{center}
1405: \end{figure}
1406: 
1407: \begin{figure}
1408:   \begin{center}
1409: \includegraphics[width=8cm,angle=0.,clip]{fig23.eps}
1410:     \caption{Configurational probabilities $P(N)$ as defined in
1411:     Sec. \ref{sec:inhom} for two non-interacting tetrahedral solutes.}
1412: \label{fig:prob_free}
1413: \end{center}
1414: \end{figure}
1415: 
1416: In the following we investigate how the explicitly resolved water
1417: affects the average orientational configurations of two close
1418: tetrahedral solutes, when their centers are held at fixed
1419: positions. In a continuous solvent the probability of observing a
1420: certain configuration is purely determined by the electrostatic
1421: interactions between the surface charges.  Obviously, for close
1422: distances of the solutes, structural effects of explicit water are
1423: expected to be very significant. A simple orientational order parameter, which coarsely
1424: probes different orientational configurations of a pair of tetrahedral
1425: solutes can be defined as follows: we count the number of surface
1426: charges in the slab delimited in width by the centers of the two
1427: solutes, as sketched in Fig.~\ref{fig:slab}. Let $N_1$ and $N_2$ be
1428: the numbers of charges in the slab belonging to the first and second
1429: solutes. The values $N_i$, $i$=1,2 for a tetrahedron with four charged
1430: vertices are obviously $N_i=1,2$ or 3 (it is not possible to have four
1431: charges on one half sphere of the tetrahedral solute). The order
1432: parameter is now defined as the product $N=N_1N_2$ and can take values
1433: 1,2,3,4,6,9, which characterizes 6 different mutual orientational
1434: configurations. For $N=1$, for instance, one charge of each solute is
1435: located in the slab, and the charges are both necessarily close to the
1436: symmetry axis; on the other hand, when $N=9$, 3 charges of each solute are
1437: within the slab and the bare triangular surfaces between the three charges are mainly
1438: facing each other. In Fig.~\ref{fig:slab} we sketch two tetrahedra
1439: in a configuration $N=3\cdot 2 = 6$.  The probability distribution
1440: $P(N)$ for two freely (without any interactions) rotating tetrahedra
1441: is plotted in Fig.~\ref{fig:prob_free}. $N=2=2\cdot 1=1\cdot 2$,
1442: $N=4=2\cdot 2$, and $N=6=2\cdot 3=3\cdot 2$ are the most likely
1443: configurations, with probabilities $P(2)\simeq0.21$, $P(4)=\simeq
1444: 0.4$, and $P(6)=0.28$.
1445: 
1446: In Fig.~\ref{fig:prob_eps80}(a)-(d) we plot the probability
1447: distribution $P(N)$ for interacting tetrahedra in a continuous solvent
1448: with permittivity $\epsilon=80$.  In Fig.~\ref{fig:prob_eps80}(a) and
1449: (b) we show the result for a pair of overall neutral $\Tn$ solutes at
1450: distances $s=3$\AA~ and $s=9$\AA~.  For the close distance the free
1451: rotator distribution is dramatically changed and the $N=1$ and $N=2$
1452: configurations are strongly favored. This is due to negative and
1453: positive charges from different solutes attracting each other at close
1454: distance. For the larger distance the electrostatic interactions are
1455: weaker and $P(N)$ strongly resembles the free rotator distribution
1456: again.  In Fig.~\ref{fig:prob_eps80}(c) and (d) we show the same
1457: distribution function, now for a pair of overall positive solutes at
1458: distances $s=3$\AA~ and $s=9$\AA~, resp. Here, at close distance the
1459: $N=1$ and $N=2$ configurations are suppressed due to the proximity of
1460: like charges, and the $N>3$ configurations are enhanced, since they
1461: allow the charges of one solute to be at larger mean distance from the
1462: like charges of the second solute. For large distances, (d), we again
1463: recover the free rotator distribution.
1464: 
1465: In Fig.~\ref{fig:prob_exp}(a)-(d) the continuous solvent is now
1466: replaced by explicit water molecules. For close solutes ($s=3$\AA) the
1467: difference with the continuous solvent is large: the probabilities of
1468: the $N=1$ and $N=2$ configurations are reduced both for the neutral
1469: (a) and overall charged tetrahedra (c). The $N=6$ and $N=9$
1470: configurations are greatly enhanced, indicating that on average the
1471: are hydrophobic surfaces face each other, as expected from the water profiles in Fig.~\ref{fig:contour_tetra}(b),(c). Remarkably,
1472: even for case (a) the explicit solvent system strongly favors water
1473: depletion rather than the proximity of two unlike charges, which would
1474: lower the electrostatic energy significantly. Increasing the distance to
1475: $s=9$\AA, the probabilities of the $N=6$ and $N=9$ configurations are
1476: lowered and the overall distributions, both for neutral and positive
1477: tetrahedra are more similar to the free rotator distribution.
1478: \begin{figure}
1479:   \begin{center}
1480: \includegraphics[width=8cm,angle=0.,clip]{fig24.eps}
1481:     \caption{Configurational probabilities $P(N)$ as defined in
1482:     Sec. \ref{sec:inhom} for two tetrahedral solutes in a continuous solvent
1483:     with $\epsilon=80$. (a) and (b) are for $\Tn$ solutes at
1484:     a surface-to-surface distance  $s=3$\AA~ and $s=9$\AA. (c)
1485:     and (d) are for $\Tp$ solutes at a distance
1486:     $s=3$\AA~ and $s=9$\AA, resp.}
1487: \label{fig:prob_eps80}
1488: \end{center}
1489: \end{figure}
1490: \begin{figure}
1491:   \begin{center}
1492: \includegraphics[width=8cm,angle=0.,clip]{fig25.eps}
1493:     \caption{Same as in Fig.~\ref{fig:prob_eps80}, but now with
1494:     explicit SPC/E water instead of a continuous solvent.}
1495: \label{fig:prob_exp}
1496: \end{center}
1497: \end{figure}
1498: \vspace{-1cm}
1499: \section{Conclusion}
1500: \label{sec:conclusion}
1501: We have used a simple model of neutral and charged, nanometer-sized
1502: spherical solutes, embedded in explicit aqueous solvent, to
1503: investigate the influence of charge patterns on the solvation of a
1504: single solute, and on the effective, solvent-induced interaction
1505: between two solutes. The charge patterns considered in this paper
1506: include uniform charge distributions (equivalent to a single charge at
1507: the center of the spherical solute), as well as tetrahedral or cubic
1508: charge distributions, involving 4 or 8 discrete positive or negative
1509: charges situated at the solute surface, adding up to an overall
1510: positive, zero or negative charge $Q$ (the $\Tp,\Tm,\Tn$ and
1511: $\Cp,\Cm,\Cn$ models).
1512:  
1513: Extensive constant pressure and constant temperature ($NPT$) Molecular
1514: Dynamics simulations were carried out under ``normal'' solvent
1515: conditions, i.e. close to liquid-vapor coexistence of water at room
1516: temperature. These simulations provide water density profiles around a
1517: single solute or a pair of solutes, which can be resolved into
1518: solute-oxygen and solute-hydrogen pair distribution functions, the
1519: distance resolved orientational order parameter $P(r)$, the solvation
1520: free energy as a function of solute radius and charge, as well as the
1521: effective force and pair potential between two solutes, averaged over
1522: solvent configurations. The main results of this investigation may be
1523: summarized as follows:
1524: 
1525: 1. The density profiles of water around a single, neutral solute
1526:    ($\Sn$ model), and their variation with solute radius $R$
1527:    (cf. Fig.~3) exhibit the characteristic ``destructuring'' for radii
1528:    $R\gtrsim 5$\AA~ already reported by earlier studies.
1529:    \cite{stilinger,huang:jpc,lum:jpc} The water molecules exhibit no
1530:    significant orientational ordering around neutral nano-sized solutes.
1531: 
1532: 2. The hydrogen and oxygen density profiles change dramatically when
1533:    the solute is uniformly charged (${\rm S}_\pm$ models).  These
1534:    profiles are sensitive to the anionic or cationic nature of the
1535:    solute (for a given absolute charge $|Q|$), in addition the
1536:    hydrogen profiles exhibit a splitting of the main peak in the case
1537:    of anionic ($\Sm$) solutes (cf. Fig.~5). The orientational order
1538:    parameter $P(r)$ exhibits a significant structure, and a relatively
1539:    slow decay with $r$, indicative of strong orientational ordering
1540:    around the solutes $\Sp$ or $\Sm$, which is somewhat more
1541:    pronounced around an anionic solute. The hydration asymmetry
1542:    results in preferential solvation of anionic solutes for a given
1543:    radius and absolute charge $|Q|$, in agreement with earlier
1544:    findings. \cite{hummer:jpc:1996,lynden-bell,garde:2004}
1545: 
1546: 3. Moving from uniformly charged solutes to discrete (tetrahedral or
1547:    cubic) charge patterns, the hydration of nano-sized solutes is found to
1548:    exhibit a strong angular modulation associated with the hydrophilic
1549:    ``patches'' around the discrete surface charges, and hydrophobic
1550:    ``patches'' in between (cf. Fig. 6). The conflicting hydration
1551:    patterns lead to a surprising depletion of water around $\Tp$ or
1552:    $\Tm$ solutes, compared to a neutral solute $\Sn$. The solvation
1553:    free energy is found to be about 20 $\%$ lower for solutes with
1554:    discrete charge patterns compared to that of uniformly charged
1555:    solutes with the same overall charge (cf. Fig. 9).
1556: 
1557: 4. The present simulations confirm the strong hydrophobic attraction
1558:    between two neutral spherical nano-sized solutes linked to solvent
1559:    ``drying'', which was already reported earlier for similar
1560:    models. The MD results for solute radii $R\gtrsim 5$\AA~ are nearly
1561:    quantitatively reproduced by a simple calculation based on purely
1562:    macroscopic considerations, and the force at solute-solute contact
1563:    is found to scale roughly linearly with $R$.
1564: 
1565: 5. The effective attraction between neutral solutes is strongly
1566:    reduced, or turns into a repulsion, when the nano-sized solutes carry
1567:    equal, uniform charge distributions. The total force has a
1568:    repulsive electrostatic component, while examination of the water
1569:    density profiles shows that the ``drying'' is mostly
1570:    suppressed. The effective force is systematically less attractive
1571:    (or more repulsive) between pairs of cationic solutes compared to
1572:    anionic pairs (cf. Fig. 17). Turning to a pair of oppositely (but
1573:    uniformly) charged solutes, the MD simulations show that the range
1574:    of the effective attraction decreases when the absolute charge
1575:    $|Q|$ increases, again in agreement with simple macroscopic
1576:    considerations, but that the effective force at contact seems to be
1577:    independent of $Q$, and equal to the hydrophobic force between
1578:    neutral solutes; we have no explanation for this surprising
1579:    observation.
1580: 
1581: 6. The situation for discrete solute charge patterns is, not
1582:    surprisingly, more complex, due to the competition between the
1583:    resulting hydrophilic and hydrophobic ``patches'' on the solute
1584:    surface. On average, some ``drying'' of water is observed, and the
1585:    resulting mean force between solutes carrying tetrahedral or cubic
1586:    patterns is once more attractive, despite the electrostatic
1587:    repulsion (``like-charge attraction''). This effect is obviously
1588:    incompatible with crude ``implicit solvent'' models.
1589: 
1590: 7.  The complete break-down of ``implicit solvent'' models, whereby
1591: the latter is replaced by a dielectric continuum, is further
1592: illustrated by the highly coarse-grained representation of the
1593: configurational probability density of two solutes carrying discrete
1594: charge distributions, introduced in Sec.~V. The relative orientations
1595: of the surface charge patterns on the two solutes are completely
1596: different for explicit and implicit solvent models, particularly at
1597: short surface-to-surface distances $s$ (cf. Figs. 24 and 25).
1598: 
1599: The key message of the present work is that explicit solvent models
1600: are unavoidable for a proper description of the effective interactions
1601: between nano-sized solutes like proteins, and that the latter are extremely
1602: sensitive to the precise location of any electric charges carried by
1603: the solutes. Contrarily to effective interactions on larger colloidal
1604: scales, a generic coarse-graining strategy appears to be useless when
1605: solutes in the nanometer range are considered, and fully molecular
1606: models are required for realistic simulations.
1607: 
1608: \section{Acknowledgments}
1609: JD acknowledges the financial support of EPSRC within the Portfolio
1610: Grant RG37352. We thank A. Archer and V. Ballenegger for useful
1611: discussions, and A.A. Louis for using the computer cluster Ice.
1612: 
1613: \section{Appendix A: Simulation details}
1614: The simulations were performed with the DLPOLY2 \cite{dlpoly}
1615: package. The Berendsen barostat and thermostat \cite{berendsen:jcp}
1616: were used to maintain the SPC/E water at a pressure of 1 bar and a
1617: temperature $T=300$K.  For the simulations of the solutes with
1618: inhomogeneous charge distributions (T and C models) we used the rigid
1619: body algorithm with quaternions to properly account for the rotation
1620: of the anisotropic solutes. To this end we switched to an integration
1621: routine using the Nos{\'e}-Hoover barostat and thermostat, which
1622: turned out to be more stable in conjunction with quaternions. We
1623: carefully checked that both barostats give the same results by
1624: performing tests with bulk water, treated both with bond contraints
1625: and with the rigid body algorithm. Test runs using the Nos{\'e}-Hoover
1626: barostat and thermostat for the S models also showed no difference.
1627: 
1628: The simulation cell is a periodically repeated cube with a maximum
1629: boxlength of about $L=48$\AA, containing up to $N_{\rm w}$=3000 water
1630: molecules, depending on the solute size. For simulations of one
1631: isolated solute we required that the surface-to-surface distance to
1632: the nearest image solute was 20\AA, yielding a box size of
1633: $L=2R+20$\AA. For the calculation of the interaction force between two
1634: solutes, the latter are placed at fixed positions ${{\vec R}_1}$ and
1635: ${{\vec R}_2}$ on the body diagonal of the simulation cell. The center
1636: to center distance is then ${R_{12}}={|{\vec R}_1}-{{\vec
1637: R}_2}|$.  The corresponding box dimensions are chosen such that the
1638: surface-to-surface distance $s={R_{12}}-2R_{0}$ to the nearest image
1639: solute is 20\AA. The box length can be calculated as $L=(4R+s+20{\rm
1640: \AA})/\sqrt{3}$.  Due to the constant pressure constraint the box length
1641: fluctuates slightly in the simulations. The long range electrostatic
1642: interactions were evaluated with smooth particle-mesh Ewald (SPME)
1643: summations \cite{essmann:jcp} using 16 $\vec k$ vectors in each
1644: direction and a convergence parameter of 3.2$/r_{\rm cut}$. A
1645: cutoff distance $r_{\rm cut}=9$\AA~ was used for LJ-interactions and
1646: the real space SPME contributions. For the nanosized solutes
1647: a larger cut-off radius is obviously required. We optimize the computational
1648: speed by introducing a second cutoff for the solute-water and
1649: solute-ion interactions, chosen to be $R_0+4$\AA, sufficient large for
1650: the shifted, short ranged repulsive interaction~(\ref{eq:solutepot}).
1651: 
1652: \section{Appendix B: Finite corrections for solvation free energies from simulation}
1653: Accurate solavtion free energies for charged
1654: solutes can be obtained by Ewald summations in periodic cells\cite{hansen} when the
1655: self-interacion energy of the solute charges with its periodic images
1656: and the background charge is properly included.\cite{hummer:jpc:1996}
1657: This correction is slightly modified when the solute size $R$ is
1658: comparable to the box size $L$.\cite{hummer:jcp:1997} The final
1659: expression for the electrostatic contribution to the solvation free
1660: energy for our solute models, including the finite size corrections,
1661: is:
1662: \begin{eqnarray}
1663: \Delta\mu_\pm &=& \Delta\mu_\pm^{\rm sim} \cr & + &
1664:  \frac{e^2}{8\pi\epsilon_0}\frac{\epsilon-1}{\epsilon} \left(\frac{\xi_{\rm
1665:  EW}}{L}\sum_{\alpha=1}^{N_{\rm c}}q_{\alpha}^{2}+\frac{2\pi
1666:  R^2}{3L^3}\sum_{\alpha=1}^{N_{\rm c}}q_{\alpha}^2\right) \cr
1667:  &+&\frac{e^2}{8\pi\epsilon_0}\frac{\epsilon-1}{\epsilon}\sum_{\alpha=1}^{N_{\rm
1668:  c}}\sum_{\beta\neq\alpha,1}^{N_{\rm c}}q_{\alpha} q_{\beta} \left(\phi_{\rm
1669:  EW}(\vec r_{\alpha\beta})-\frac{1}{|\vec r_{\alpha\beta}|}\right),\nonumber \\
1670: \label{eq:correct}
1671: \end{eqnarray}
1672: where $\epsilon$ is the macroscopic permittivity of water, and for a
1673: periodic array of cubic simulation cells,
1674: $\xi_{EW}\approx-2.837297$. In the case of a uniformly charged
1675: solute corresponding to a single charged site $qe$ at the center, the
1676: result (\ref{eq:correct}) reduces to\cite{hummer:jcp:1997}
1677: \begin{eqnarray}
1678: \Delta\mu_{\pm}=\Delta\mu_\pm^{\rm sim}+\frac{q^2 e^2}{8\pi\epsilon_0}\frac{\epsilon-1}{\epsilon}\left(\frac{\xi_{\rm EW}}{L}+\frac{2\pi R^2}{3L^3}\right).
1679: \label{eq:correct2}
1680: \end{eqnarray}
1681: With the system sizes used in the present simulation the finite size
1682: corrections are very large and represent typically twice the value of
1683: $\Delta\mu_{\rm sim}$ and stem mainly from the $\xi_{\rm EW}$-term in
1684: Eqs.~(19) and (20).
1685: 
1686: \begin{thebibliography}{29}
1687: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1688: \expandafter\ifx\csname bibnamefont\endcsname\relax
1689:   \def\bibnamefont#1{#1}\fi
1690: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1691:   \def\bibfnamefont#1{#1}\fi
1692: \expandafter\ifx\csname citenamefont\endcsname\relax
1693:   \def\citenamefont#1{#1}\fi
1694: \expandafter\ifx\csname url\endcsname\relax
1695:   \def\url#1{\texttt{#1}}\fi
1696: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1697: \providecommand{\bibinfo}[2]{#2}
1698: \providecommand{\eprint}[2][]{\url{#2}}
1699: 
1700: \bibitem[{\citenamefont{Rubinstein and Colby}(2003)}]{rubinstein}
1701: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Rubinstein}} \bibnamefont{and}
1702:   \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Colby}},
1703:   \emph{\bibinfo{title}{Polymer Physics}} (\bibinfo{publisher}{Oxford
1704:   University Press}, \bibinfo{year}{2003}).
1705: 
1706: \bibitem[{\citenamefont{Chandler}(2004)}]{chandler_review}
1707: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}}
1708:   (\bibinfo{year}{2004}), \bibinfo{note}{to appear in Nature}.
1709: 
1710: \bibitem[{\citenamefont{Likos}(2001)}]{likos:physrep}
1711: \bibinfo{author}{\bibfnamefont{C.~N.} \bibnamefont{Likos}},
1712:   \bibinfo{journal}{Phys. Rep.} \textbf{\bibinfo{volume}{348}},
1713:   \bibinfo{pages}{267} (\bibinfo{year}{2001}).
1714: 
1715: \bibitem[{\citenamefont{Stilinger}(1973)}]{stilinger}
1716: \bibinfo{author}{\bibfnamefont{F.~H.} \bibnamefont{Stilinger}},
1717:   \bibinfo{journal}{J. Solution Chem.} \textbf{\bibinfo{volume}{2}},
1718:   \bibinfo{pages}{141} (\bibinfo{year}{1973}).
1719: 
1720: \bibitem[{\citenamefont{Hummer and Garde}(1998)}]{hummer:prl}
1721: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}} \bibnamefont{and}
1722:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Garde}},
1723:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{80}},
1724:   \bibinfo{pages}{4193} (\bibinfo{year}{1998}).
1725: 
1726: \bibitem[{\citenamefont{Lum et~al.}(1999)\citenamefont{Lum, Chandler, and
1727:   Weeks}}]{lum:jpc}
1728: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lum}},
1729:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}}, \bibnamefont{and}
1730:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Weeks}},
1731:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{103}},
1732:   \bibinfo{pages}{4570} (\bibinfo{year}{1999}).
1733: 
1734: \bibitem[{\citenamefont{Huang et~al.}(2001)\citenamefont{Huang, Geissler, and
1735:   Chandler}}]{huang:jpc}
1736: \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{Huang}},
1737:   \bibinfo{author}{\bibfnamefont{P.~L.} \bibnamefont{Geissler}},
1738:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}},
1739:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{105}},
1740:   \bibinfo{pages}{6704} (\bibinfo{year}{2001}).
1741: 
1742: \bibitem[{\citenamefont{Wallquist and Berne}(1995)}]{wallquist:jpc}
1743: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Wallquist}} \bibnamefont{and}
1744:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Berne}},
1745:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{99}},
1746:   \bibinfo{pages}{2893} (\bibinfo{year}{1995}).
1747: 
1748: \bibitem[{\citenamefont{Shinto et~al.}(1999)\citenamefont{Shinto, Miyahara, and
1749:   Higashitani}}]{shinto}
1750: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Shinto}},
1751:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Miyahara}}, \bibnamefont{and}
1752:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Higashitani}},
1753:   \bibinfo{journal}{J. Coll. Interface Sci.} \textbf{\bibinfo{volume}{209}},
1754:   \bibinfo{pages}{79} (\bibinfo{year}{1999}).
1755: 
1756: \bibitem[{\citenamefont{Kinoshita et~al.}(1996)\citenamefont{Kinoshita, Iba,
1757:   Kuwamoto, and Harada}}]{kinoshita}
1758: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kinoshita}},
1759:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Iba}},
1760:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kuwamoto}}, \bibnamefont{and}
1761:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Harada}},
1762:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{105}},
1763:   \bibinfo{pages}{7177} (\bibinfo{year}{1996}).
1764: 
1765: \bibitem[{\citenamefont{Qin and Fichthorn}(2003)}]{qin}
1766: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Qin}} \bibnamefont{and}
1767:   \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Fichthorn}},
1768:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{119}},
1769:   \bibinfo{pages}{9745} (\bibinfo{year}{2003}).
1770: 
1771: \bibitem[{\citenamefont{Allahyarov et~al.}(2002)\citenamefont{Allahyarov,
1772:   L{\"o}wen, Louis, and Hansen}}]{elshad:epl:2002}
1773: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Allahyarov}},
1774:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{L{\"o}wen}},
1775:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Louis}}, \bibnamefont{and}
1776:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hansen}},
1777:   \bibinfo{journal}{Europhys. Lett.} \textbf{\bibinfo{volume}{57}},
1778:   \bibinfo{pages}{731} (\bibinfo{year}{2002}).
1779: 
1780: \bibitem[{\citenamefont{Berendsen et~al.}(1987)\citenamefont{Berendsen,
1781:   Grigera, and Straatsma}}]{berendsen:jpc}
1782: \bibinfo{author}{\bibfnamefont{H.~J.~C.} \bibnamefont{Berendsen}},
1783:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Grigera}},
1784:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.~P.}
1785:   \bibnamefont{Straatsma}}, \bibinfo{journal}{J. Phys. Chem.}
1786:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{6269} (\bibinfo{year}{1987}).
1787: 
1788: \bibitem[{\citenamefont{Dzubiella and Hansen}(2003)}]{dzubiella:jcp:2003}
1789: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Dzubiella}} \bibnamefont{and}
1790:   \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1791:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{119}},
1792:   \bibinfo{pages}{12049} (\bibinfo{year}{2003}).
1793: 
1794: \bibitem[{\citenamefont{Spohr}(1999)}]{spohr:1999}
1795: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Spohr}},
1796:   \bibinfo{journal}{Electrochim. Acta} \textbf{\bibinfo{volume}{44}},
1797:   \bibinfo{pages}{1697} (\bibinfo{year}{1999}).
1798: 
1799: \bibitem[{\citenamefont{Essmann et~al.}(1995)\citenamefont{Essmann, Perera,
1800:   Berkowitz, Darden, Lee, and Pedersen}}]{essmann:jcp}
1801: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Essmann}},
1802:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Perera}},
1803:   \bibinfo{author}{\bibfnamefont{M.~L.} \bibnamefont{Berkowitz}},
1804:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Darden}},
1805:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Lee}}, \bibnamefont{and}
1806:   \bibinfo{author}{\bibfnamefont{L.~G.} \bibnamefont{Pedersen}},
1807:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{103}},
1808:   \bibinfo{pages}{8577} (\bibinfo{year}{1995}).
1809: 
1810: \bibitem[{\citenamefont{Smith and Forester}(1999)}]{dlpoly}
1811: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Smith}} \bibnamefont{and}
1812:   \bibinfo{author}{\bibfnamefont{T.~R.} \bibnamefont{Forester}}
1813:   (\bibinfo{year}{1999}), \bibinfo{note}{the DLPOLY\_2 User Manual}.
1814: 
1815: \bibitem[{\citenamefont{Frenkel and Smit}(1996)}]{frenkelsmit}
1816: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Frenkel}} \bibnamefont{and}
1817:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Smit}},
1818:   \emph{\bibinfo{title}{Understanding Molecular Simulation: From Algorithms to
1819:   Applications}} (\bibinfo{publisher}{Academic Press}, \bibinfo{year}{1996}).
1820: 
1821: \bibitem[{\citenamefont{Born}(1920)}]{born}
1822: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Born}}, \bibinfo{journal}{Z.
1823:   Phys.} \textbf{\bibinfo{volume}{1}}, \bibinfo{pages}{45}
1824:   (\bibinfo{year}{1920}).
1825: 
1826: \bibitem[{\citenamefont{Reiss et~al.}(1959)\citenamefont{Reiss, Frisch, and
1827:   Lebowitz}}]{reiss:jcp:1959}
1828: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Reiss}},
1829:   \bibinfo{author}{\bibfnamefont{H.~L.} \bibnamefont{Frisch}},
1830:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~L.}
1831:   \bibnamefont{Lebowitz}}, \bibinfo{journal}{J. Chem. Phys.}
1832:   \textbf{\bibinfo{volume}{31}}, \bibinfo{pages}{369} (\bibinfo{year}{1959}).
1833: 
1834: \bibitem[{\citenamefont{Latimer et~al.}(1939)\citenamefont{Latimer, Pitzer, and
1835:   Slansky}}]{latimer:jcp:1939}
1836: \bibinfo{author}{\bibfnamefont{W.~M.} \bibnamefont{Latimer}},
1837:   \bibinfo{author}{\bibfnamefont{K.~S.} \bibnamefont{Pitzer}},
1838:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.~M.}
1839:   \bibnamefont{Slansky}}, \bibinfo{journal}{J. Chem. Phys.}
1840:   \textbf{\bibinfo{volume}{7}}, \bibinfo{pages}{108} (\bibinfo{year}{1939}).
1841: 
1842: \bibitem[{\citenamefont{Hummer et~al.}(1996)\citenamefont{Hummer, Pratt, and
1843:   Garcia}}]{hummer:jpc:1996}
1844: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1845:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Pratt}}, \bibnamefont{and}
1846:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1847:   \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{100}},
1848:   \bibinfo{pages}{1206} (\bibinfo{year}{1996}).
1849: 
1850: \bibitem[{\citenamefont{Hummer et~al.}(1997)\citenamefont{Hummer, Pratt, and
1851:   Garcia}}]{hummer:jcp:1997}
1852: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hummer}},
1853:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Pratt}}, \bibnamefont{and}
1854:   \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Garcia}},
1855:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{107}},
1856:   \bibinfo{pages}{9275} (\bibinfo{year}{1997}).
1857: 
1858: \bibitem[{\citenamefont{Lynden-Bell and Rasaiah}(1997)}]{lynden-bell}
1859: \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{Lynden-Bell}}
1860:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~C.}
1861:   \bibnamefont{Rasaiah}}, \bibinfo{journal}{J. Chem. Phys.}
1862:   \textbf{\bibinfo{volume}{107}}, \bibinfo{pages}{1981} (\bibinfo{year}{1997}).
1863: 
1864: \bibitem[{\citenamefont{Rajamani et~al.}(2004)\citenamefont{Rajamani, Ghosh,
1865:   and Garde}}]{garde:2004}
1866: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Rajamani}},
1867:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Ghosh}}, \bibnamefont{and}
1868:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Garde}}, \bibinfo{journal}{J.
1869:   Chem. Phys.} \textbf{\bibinfo{volume}{120}}, \bibinfo{pages}{4457}
1870:   (\bibinfo{year}{2004}).
1871: 
1872: \bibitem[{\citenamefont{Bolhuis and Chandler}(2000)}]{bolhuis}
1873: \bibinfo{author}{\bibfnamefont{P.~G.} \bibnamefont{Bolhuis}} \bibnamefont{and}
1874:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chandler}},
1875:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{113}},
1876:   \bibinfo{pages}{8154} (\bibinfo{year}{2000}).
1877: 
1878: \bibitem[{\citenamefont{Louis et~al.}(2002)\citenamefont{Louis, Bolhuis, and
1879:   Hansen}}]{louis:jcp}
1880: \bibinfo{author}{\bibfnamefont{A.~A.} \bibnamefont{Louis}},
1881:   \bibinfo{author}{\bibfnamefont{P.~G.} \bibnamefont{Bolhuis}},
1882:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~P.}
1883:   \bibnamefont{Hansen}}, \bibinfo{journal}{J. Chem. Phys.}
1884:   \textbf{\bibinfo{volume}{117}}, \bibinfo{pages}{1893} (\bibinfo{year}{2002}).
1885: 
1886: \bibitem[{\citenamefont{Berendsen et~al.}(1984)\citenamefont{Berendsen, Postma,
1887:   van Gunsteren, DiNola, and Haak}}]{berendsen:jcp}
1888: \bibinfo{author}{\bibfnamefont{H.~J.~C.} \bibnamefont{Berendsen}},
1889:   \bibinfo{author}{\bibfnamefont{J.~P.~M.} \bibnamefont{Postma}},
1890:   \bibinfo{author}{\bibfnamefont{W.~F.} \bibnamefont{van Gunsteren}},
1891:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{DiNola}}, \bibnamefont{and}
1892:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Haak}},
1893:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{81}},
1894:   \bibinfo{pages}{3684} (\bibinfo{year}{1984}).
1895: 
1896: \bibitem[{\citenamefont{Hansen}(1986)}]{hansen}
1897: \bibinfo{author}{\bibfnamefont{J.-P.} \bibnamefont{Hansen}},
1898:   \emph{\bibinfo{title}{Molecular Dynamics Simulation of Statistical Mechanical
1899:   Systems}} (\bibinfo{publisher}{North Holland}, \bibinfo{address}{Amsterdam},
1900:   \bibinfo{year}{1986}), \bibinfo{note}{edited by G. Ciccotti and W.G. Hoover}.
1901: 
1902: \end{thebibliography}
1903: 
1904: 
1905: \end{document}
1906: 
1907: 
1908: 
1909: 
1910: 
1911: 
1912: