1:
2: %\documentclass[byrevtex,prb,superscriptaddress,showpacs,showkeys,twocolumn]{revtex4}
3: \documentclass[byrevtex,pre,superscriptaddress,showpacs,showkeys,preprint,nofootinbib]{revtex4}
4: \usepackage{amsmath,amssymb}
5: \usepackage{epsfig}
6:
7: \newcommand{\epl}{Europhys. Lett.}
8:
9: \newcommand{\rcite}[1]{Ref.~\onlinecite{#1}}
10: \newcommand{\rcites}[1]{Refs.~\onlinecite{#1}}
11:
12: \def\br{{\bf r}}
13: \def\bn{{\bf n}}
14: %\def\en{{\bf e}_n}
15: %\def\et{{\bf e}_t}
16: \def\ez{{\bf e}_z}
17: \def\er{{\bf e}_r}
18: \def\epsf{\varepsilon_F}
19: \def\epsp{\varepsilon_\Pi}
20: \def\epsm{\varepsilon_{\Pi_m}}
21: \def\hepsf{\hat{\varepsilon}_F}
22: \def\hepsp{\hat{\varepsilon}_\Pi}
23: \def\hsmen{\hat{S}_{\rm men}}
24: \def\vmen{V_{\rm men}}
25: \def\avsup{\mbox{}^aV_{\rm sup}}
26: \def\avcorr{\mbox{}^aV_{\rm corr}}
27: \def\bvsup{\mbox{}^bV_{\rm sup}}
28: \def\bvcorr{\mbox{}^bV_{\rm corr}}
29: %
30: \newcommand{\Ezp}{\hat{E}_{z,+}}
31: \newcommand{\Ezm}{\hat{E}_{z,-}}
32: \newcommand{\Ep}{\hat{\bf E}_{\parallel}}
33: %\newcommand{\Ep}{E_{\parallel,0}}
34:
35: \begin{document}
36:
37:
38: \title{Theory of capillary--induced interactions beyond the
39: superposition approximation}
40:
41: \author{Alvaro Dom\'\i nguez}
42: \affiliation{F\'\i sica Te\'orica, Universidad de Sevilla, Apdo.\ 1065, E--41080 Sevilla, Spain}
43: \author{Martin Oettel}
44: \affiliation{Institut f\"ur Physik, Universit{\"a}t Mainz, WA 331, D-55099 Mainz, Germany}
45: \author{S.\ Dietrich}
46: \affiliation{Max--Planck--Institut f\"ur Metallforschung, Heisenbergstr.\ 3, D--70569 Stuttgart, Germany}
47: \affiliation{Institut f\"ur Theoretische und Angewandte
48: Physik, Universit\"at Stuttgart, Pfaffenwaldring 57, D--70569 Stuttgart, Germany}
49:
50: \date{June 1, 2007}
51:
52: \begin{abstract}
53: Within a general theoretical framework we study the effective,
54: deformation--induced interaction between two colloidal particles
55: trapped at a fluid interface in the regime of small
56: deformations. In many studies, this interaction has been computed
57: with the ansatz that the actual interface configuration for the pair
58: is given by the linear superposition of the interface deformations
59: around the single particles.
60: Here we assess the validity of this approach and compute the
61: leading term of the effective interaction for large interparticle separation
62: beyond this so-called superposition
63: approximation. As an application, we consider the experimentally
64: relevant case of interface deformations owing to the electrostatic
65: field emanating from charged colloidal particles. In mechanical
66: isolation, i.e., if the net force acting on the total system
67: consisting of the particles plus the interface vanishes, the
68: superposition approximation is actually invalid. The effective
69: capillary interaction is governed by contributions beyond this
70: approximation and turns out to be attractive. For sufficiently small
71: surface charges on the colloids, such that linearization is strictly
72: valid, and at asymptotically large separations, the effective
73: interaction does not overcome the direct electrostatic repulsion
74: between the colloidal particles.
75: \end{abstract}
76:
77: \pacs{82.70.Dd; 68.03.Cd}
78: \keywords{Colloids; surface tension and related phenomena}
79:
80: \maketitle
81:
82: \section{Introduction}
83: \label{sec:intro}
84:
85: The self--assembly of sub-$\mu$m colloidal particles at fluid (e.g.,
86: water/air or water/oil) interfaces has gained significant interest in
87: view of various basic and applied issues such as the study of
88: two--dimensional melting \cite{Pier80}, investigations of mesoscale
89: structure formation \cite{Joan01}, and engineering of colloidal
90: crystals on spherical surfaces \cite{DHNM02}.
91: %
92: The colloidal particles are trapped at the interface if the fluid
93: phases wet the colloid only partially; this configuration is stable
94: against thermal fluctuations and it appears to be even the global
95: equilibrium state, in accordance with the experimental observation that the
96: colloids immersed in the bulk phases are attracted towards the
97: interface \cite{Pier80}.
98:
99: In order to prevent coagulation, the colloidal particles are
100: electrically charged. The ensuing repulsive force is well understood
101: and at large separations $d$ it varies like a dipole--dipole
102: interaction $\propto 1/d^4$, because the monopoles vanish due to
103: screening by counterions in water \cite{Hurd85,ACNP00}.
104: %
105: Nevertheless, several experimental findings have led to postulating an
106: attractive effective force between such particles with a range much
107: larger than that of van--der--Waals forces
108: \cite{RGI97,GhEa97,SDJ00,QMMH01,GEFM01,MGIR02,NBHD02,TAKK03,GoRu05,CTHN06} (but see also \rcite{FMMH04}).
109: Spherical particles (radii $R=0.25\dots 2.5$ $\mu$m) at flat
110: water--air interfaces exhibit the spontaneous formation of complicated
111: metastable mesostructures consistent with the presence of a minimum in
112: the effective intercolloidal potential at separations $d/R\approx
113: 3\dots 20$ and with a depth of at least a few $k_B T$.
114:
115: Until now, no unequivocal explanation for the appearance of these
116: relatively long--ranged attractions is available. One possibility,
117: which has been explored intensively in previous years, consists of an
118: attraction mediated by the deformation of the interface (see
119: Fig.~\ref{fig:2coll_side}). This is similar to the so-called flotation
120: force attracting particles floating at the surface of water which is
121: deformed by the weight of the particles \cite{Nico49,CHW81}. But
122: gravity plays no role for micrometer sized particles as described
123: above. Instead, the electrostatic field around the charged colloids
124: deforms the interface and gives rise to effective, capillary--induced
125: interactions.
126: %
127: There have been contradictory results about the properties of this
128: effective interaction. In \rcites{NBHD02,DKB04} it has been
129: argued in favor
130: of an attractive force decaying like $1/d$ (similar to the
131: gravity--induced flotation force). In
132: \rcites{MeAi03,FoWu04,ODD05a,DOD05,ODD06} this line of thought
133: was shown to be invalid because the decay
134: turns out to be
135: much faster if the physical system consisting of the colloidal
136: particles plus the interface is mechanically isolated, i.e., if the
137: total force acting on this system vanishes in the limit of a
138: macroscopically extended interface with negligible borders. This is in
139: principle the case for the experiments conducted in Langmuir troughs
140: with lateral extensions several times the capillary length of the
141: interface (i.e., orders of magnitude larger than the colloid radius).
142: In this case the electrostatic force pulling on the interface is
143: counterbalanced exactly by the electrostatic force pushing the
144: particles into water.
145: %
146: In particular, the authors of \rcite{MeAi03} argue for an attractive
147: force decaying like $1/d^7$. However, this was in turn corrected in
148: \rcites{FoWu04,ODD05a} and a repulsive force with an asymptotic
149: decay $\propto 1/d^7$ was derived. In \rcite{ODD05a} it was noticed
150: that this latter result is actually unreliable because the linear
151: superposition approximation employed in these calculations is not valid
152: if mechanical isolation holds. Furthermore, since it is the electric
153: field ${\bf E}$, and not the electrostatic pressure $\propto {\bf
154: E}^2$, which obeys a superposition principle, the superposition
155: approximation will be unsuitable if the main contribution to the
156: electrostatic pressure stems from cross terms in ${\bf E}^2$. This
157: latter case was studied in \rcites{ODD05b,WuFo05} with the
158: conclusion that the capillary--induced force is attractive and decays
159: like $1/d^4$, which is the same asymptotic behavior as the direct
160: dipole--dipole repulsion; whether the capillary attraction overcomes
161: the electric repulsion must then be determined by a detailed analysis
162: of the electrostatic problem. The total force is repulsive in the
163: regime of small deformations of a flat interface (equivalent to the
164: regime of small colloidal charges), for which the calculations were
165: carried out \cite{ODD05b,WuFo05}. Only for sufficiently large
166: colloidal charges the capillary attraction may asymptotically overcome
167: the electrostatic repulsion leading to a minimum in the total
168: effective potential \cite{ODD05b}. However, further calculations ---
169: going beyond the linearization assumptions in treating the
170: electrostatic pressure on the interface and the energy of the deformed
171: interface --- are necessary to substantiate this claim.
172: %
173: The experiment described in \rcite{NBHD02} is peculiar in the sense that
174: the unperturbed interface is actually that of a relatively small
175: spherical droplet pending from a plate. The importance of this
176: finite--size effect was studied in
177: \rcites{ODD05a,DOD05,DOD06b,Wuer06b,DOD07a}. In
178: \rcites{DOD05,DOD06b,DOD07a} it was found that the flotation--like
179: decay $1/d$ is present because the plate breaks the condition of
180: mechanical isolation, but it is quantitatively too small to explain
181: the experimental observations\footnote{In this respect, \rcite{ODD05a}
182: is incomplete but the same conclusion about the
183: irrelevance of this effect is reached. \rcite{Wuer06b} contains
184: several important mathematical errors \cite{DOD07a}
185: and the corresponding conclusion erroneously disagrees with the
186: actual irrelevance of the finite--size effect.}.
187:
188: \begin{figure}
189: \begin{center}
190: \epsfig{file=figures/2coll_side_electric.eps,width=.5\textwidth}
191: \caption{ Schematic drawing of the deformation of a fluid
192: interface by electric fields due to charged colloidal particles
193: trapped at the interface. Counterions gather on the side of the
194: electrolytic phase (water) and a pressure field arises which
195: pulls on the interface and on the particles.}
196: \label{fig:2coll_side}
197: \end{center}
198: \end{figure}
199:
200:
201: Here we follow the approach of \rcite{ODD05a} in order to calculate
202: the capillary--induced effective interaction beyond the superposition
203: approximation in the case that there is a pressure field of general form
204: $\hat{\Pi}(\br)$ acting on the interface in the limit of small
205: deformations. The limit of small deformations corresponds to an
206: analysis to leading order in the small dimensionless parameters
207: $\hepsf$ and $\hepsp$, defined in Eq.~(\ref{eq:epsfp}) below,
208: which are measures of the force
209: acting on the colloidal particles and the interface, respectively.
210: In this manner, inter alia
211: we provide a mathematically sound derivation of the results reported
212: in \rcites{ODD05b,WuFo05}. In Sec.~\ref{sec:theory} we derive the
213: exact leading--order expressions for the capillary potential and show
214: in which respect the superposition approximation becomes inconsistent
215: if mechanical isolation holds. In such a case, the final result
216: depends
217: on the stage at which the approximation is introduced and applied. In
218: Subsec.~\ref{sec:asymptotic} we derive the leading asymptotic dependence
219: on $d$ of the effective interaction using only some rather general
220: assumptions on the form of $\hat{\Pi}(\br)$. The result for the
221: effective potential energy is summarized in Eq.~(\ref{eq:vmen_asymp}).
222: In Sec.~\ref{sec:appl} we analyze the particular case that the
223: pressure field $\hat{\Pi}(\br)$ is due to the electric field created
224: by charged colloidal particles in a mechanically isolated system. We
225: consider two limiting cases in order to solve the electrostatic
226: problem: (i) water as a perfect conductor (i.e., vanishing Debye
227: length), and (ii) the colloidal particles as point--like objects (i.e.,
228: vanishing particle radius). In both cases we recover the conclusion
229: that the effective interaction is attractive but can overcome the
230: electric repulsion only if $\hepsf\gtrsim 1$, which is outside the
231: small--deformation regime considered here.
232: %
233: Finally, in Sec.~\ref{sec:end} we summarize our results and discuss
234: their relevance in connection with the experiments described in the
235: literature.
236:
237:
238: \section{Free energy of effective capillary interaction}
239: \label{sec:theory}
240:
241: \subsection{Exact results}
242: \label{sec:exact}
243:
244: We consider $N$ identical\footnote{In actual experiments there is a
245: certain degree of polydispersity in size, shape, electric charge,
246: etc., which we do not expect to alter the conclusions qualitatively
247: if it remains small enough.} spherical particles trapped at a fluid
248: interface (see Fig.~\ref{fig:2coll_top} for $N=2$). We define the
249: reference configuration as a flat interface in the plane $z=0$ and the
250: colloids at a height such that the colloid--interface contact occurs
251: at Young's contact angle $\theta \in (0,\pi)$. In this configuration,
252: $S_\alpha$ denotes the circular disk delimited by the contact line on
253: the colloid $\alpha$, $\partial S_\alpha$ is the corresponding contact
254: line (of radius $r_0 = R \sin\theta$ for a particle of radius $R$)
255: traced counterclockwise when viewed from the
256: top\footnote{Which fluid phase is to be the top one depends on the
257: experimental setup. Usually gravity breaks the up--down
258: symmetry.}, and $\hsmen$
259: is the fluid interface (with surface tension $\gamma$), enclosed by a
260: boundary $C_L$ of typical size of the order of $L$ representing, e.g.,
261: the vessel containing the system.
262: The relative lateral positions of the colloids are kept fixed and thus we
263: consider only vertical displacements $\hat{u}(\br=(x,y))$ of the fluid
264: interface and of the height $\Delta \hat{h}_\alpha$ of the center of
265: the colloid $\alpha$ relative to the plane $z=0$. In the reference
266: configuration there is a (vertical) force $\hat{F}_\alpha$ acting on
267: the colloid $\alpha$ and a (vertical) force per unit area
268: $\hat{\Pi}(\br)$ on the meniscus. We define the dimensionless forces
269: \begin{equation}
270: \label{eq:epsfp}
271: \hat{\varepsilon}_{{F}_\alpha} := - \frac{\hat{F}_\alpha}{2 \pi \gamma r_0} , \qquad
272: \hepsp := \frac{1}{2 \pi \gamma r_0}
273: \int_{\hsmen} \!\!\!\!\! dA \; \hat{\Pi} ,
274: \end{equation}
275: where $dA$ is the element of interface area.
276: %
277: The reference configuration is the equilibrium state in the absence of
278: forces, $\hat{F}_\alpha \equiv 0, \hat{\Pi} \equiv 0$. Within the
279: approximation of small deviations from the reference configuration,
280: the free energy of the system with respect to this configuration is
281: \cite{ODD05a}
282: \begin{equation}
283: \label{eq:freeF}
284: {\cal \hat{F}} =
285: \gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
286: \left[ \frac{1}{2} |\nabla \hat{u}|^2
287: - \frac{1}{\gamma} \hat{\Pi} \, \hat{u} \right] + \sum_{\alpha=1}^N \left\{
288: \frac{\gamma}{2 r_{0}} \oint_{\partial S_\alpha} \!\!\! d\ell \; [\Delta \hat{h}_\alpha - \hat{u}]^2
289: - \hat{F}_\alpha \Delta \hat{h}_\alpha \right\}
290: + {\cal O}(\hepsf,\hepsp)^3,
291: \end{equation}
292: where $d\ell$ is the element of arclength.
293: %
294: The free energy contains a contribution from the change of the contact
295: area between the phases (two fluid phases and the solid particles),
296: and a contribution from the work done by the forces $\hat{F}_\alpha$
297: and $\hat{\Pi}$ via displacements from the reference configuration.
298:
299: \begin{figure}
300: \begin{center}
301: \epsfig{file=figures/2coll_top.eps,width=.6\textwidth}
302: \caption{ Top view (plane $z=0$) of the reference configuration
303: with two colloids. $d$ is the fixed lateral distance between
304: the colloid centers projected onto the plane $z=0$.
305: $S_1$ and $S_2$ are disks
306: of radius $r_{0}$, the corresponding circumferences
307: (counterclockwise) are $\partial S_1$ and $\partial S_2$. The
308: (projected) interface is
309: $\hsmen = \mathbb{R}^2\backslash(S_1 \bigcup S_2)$.
310: The position of any point on the plane is denoted with $\br$; in
311: particular, $\br_\alpha$ is the position of the
312: colloid $\alpha$.
313: }
314: \label{fig:2coll_top}
315: \end{center}
316: \end{figure}
317:
318: The values of $\hat{u}({\bf r})$ and $\Delta \hat{h}_\alpha$ in the
319: equilibrium state are determined by minimizing this free energy. This
320: leads to the following equations:
321: \begin{subequations}
322: \label{eq:equil}
323: \begin{align}
324: \label{eq:equil_h}
325: \Delta \hat{h}_\alpha &=
326: \langle{\hat{u}}\rangle_\alpha -
327: r_0 \, \hat{\varepsilon}_{{F}_\alpha} ,
328: & &\langle{\cdot}\rangle_\alpha := \frac{1}{2 \pi r_{0}}
329: \oint_{\partial S_\alpha} \!\!\! d\ell \; (\cdot) , \\
330: \label{eq:equil_YL}
331: \nabla^2 \hat{u} &=
332: - \frac{1}{\gamma} \hat{\Pi} ,
333: & &\br \in \hsmen \\
334: \label{eq:equil_bc}
335: \bn_\alpha \cdot \nabla \hat{u} ({\bf r}) &=
336: \hat{\varepsilon}_{{F}_\alpha} +
337: \frac{\hat{u}({\bf r})-\langle{\hat{u}}\rangle_\alpha}{r_{0}} ,
338: & &\br \in \partial S_\alpha , \\
339: \label{eq:equil_pin}
340: \hat{u}({\bf r}) &= 0 , & &\br \in C_L ,
341: \end{align}
342: \end{subequations}
343: where $\bn_\alpha$ is the unit vector in the outward
344: normal direction of $\partial S_\alpha$.
345: Equation~(\ref{eq:equil_h}) is a geometrical relationship,
346: Eq.~(\ref{eq:equil_YL}) describes local mechanical equilibrium (the
347: pressure $\hat{\Pi}$ is compensated by the curvature--induced
348: interfacial tension), Eq.~(\ref{eq:equil_bc}) describes mechanical
349: equilibrium of the particle (the force $\hat{F}_\alpha$ is balanced by
350: the interfacial tension exerted at the contact line), and
351: %
352: Eq.~(\ref{eq:equil_pin}) represents a boundary condition at the
353: external border $C_L$ (for simplicity we take a pinned interface, but
354: of course other physically reasonable boundary conditions are
355: possible, the details of which are actually irrelevant in the limit
356: $L\to\infty$ we shall consider \cite{ODD05a}).
357: %
358: The free energy functional in Eq.~(\ref{eq:freeF}) {\em evaluated at
359: the equilibrium configuration} can be simplified by using the
360: relationships in Eq.~(\ref{eq:equil}) and we obtain two equivalent
361: expressions:
362: \begin{subequations}
363: \label{eq:equilF}
364: \begin{eqnarray}
365: \label{eq:freeF1}
366: {\cal \hat{F}}_{\rm eq} & = &
367: - \frac{1}{2}\gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
368: |\nabla \hat{u}|^2
369: + \pi \gamma \sum_{\alpha=1}^N \left[
370: \langle\hat{u}\rangle_\alpha^2 - \langle\hat{u}^2\rangle_\alpha
371: - r_0^2 \, \hat{\varepsilon}_{{F}_\alpha}^2
372: \right] \\
373: \label{eq:freeF2}
374: & = & - \frac{1}{2} \int_{\hsmen} \!\!\!\!\!\! dA \;
375: \hat{\Pi} \, \hat{u}
376: + \pi \gamma \sum_{\alpha=1}^N r_0 \, \hat{\varepsilon}_{{F}_\alpha}
377: \left[ \langle\hat{u}\rangle_\alpha
378: - r_0 \, \hat{\varepsilon}_{{F}_\alpha}
379: \right] .
380: \end{eqnarray}
381: \end{subequations}
382:
383: From this point onward, we shall consider the particular case of two
384: identical colloids. We assume that any external force acting on the
385: system (e.g., gravity) is independent of the positions of the
386: particles\footnote{This seems to be actually a reliable approximation for the
387: experimental setups considered so far in the literature, but see the
388: discussion in Sec.\ \ref{sec:end}.}, so that symmetry
389: arguments
390: will allow one to simplify the expressions.
391: %
392: We use the notations $u_\alpha ({\bf r}) := u(|{\bf r} - {\bf
393: r}_\alpha|)$, $\Pi_\alpha ({\bf r}) := \Pi(|{\bf r} - {\bf
394: r}_\alpha|)$, $\epsf := \hat{\varepsilon}_{F_\alpha}$, $S_{\rm men,
395: \alpha} := \mathbb{R}^2\backslash S_\alpha$ for the corresponding
396: quantities in the presence of a {\em single} colloid located at
397: position ${\bf r}_\alpha$. This means that the function $u_\alpha({\bf
398: r}) \; (\alpha=1, 2)$ satisfies
399: \begin{subequations}
400: \label{eq:single}
401: \begin{align}
402: \label{eq:single_a}
403: \nabla^2 {u}_\alpha &=
404: - \frac{1}{\gamma} {\Pi}_\alpha ,
405: & &\br \in {S}_{\rm men, \alpha} , \\
406: \bn_\alpha \cdot \nabla u_\alpha ({\bf r}) &=
407: \epsf,
408: & &\br \in \partial S_\alpha , \\
409: \label{eq:single_pin}
410: {u}_\alpha({\bf r}) &= 0
411: & &\br \in C_L .
412: \end{align}
413: \end{subequations}
414: In terms of these single--colloid solutions the configuration
415: in the presence of two colloids can be written without loss of
416: generality as
417: \begin{subequations}
418: \begin{align}
419: \label{eq:hpi}
420: \hat{\Pi} &= \Pi_1 + \Pi_2 + 2 \Pi_m , \\
421: \label{eq:hu}
422: \hat{u} &= u_1 + u_2 + u_m , \\
423: \label{eq:hepsf}
424: \hepsf &= \epsf + \varepsilon_m ,
425: \end{align}
426: \end{subequations}
427: where we have introduced $\hepsf := \hat{\varepsilon}_{F_1} =
428: \hat{\varepsilon}_{F_2}$ reflecting the symmetry of the
429: problem.
430: The fields $u_m(\br)$ and $\Pi_m(\br)$ and the quantity
431: $\varepsilon_m$ introduced this way represent
432: the corrections to the so-called superposition approximation, which is
433: defined by setting
434: \begin{subequations}
435: \label{eq:superpositionAnsatz}
436: \begin{eqnarray}
437: {u}_m & = & 0 , \\
438: {\Pi}_m & = & 0 , \\
439: \varepsilon_m & = & 0 ,
440: \end{eqnarray}
441: \end{subequations}
442: i.e., the effects of other particles on the single--particle
443: configuration are neglected altogether.
444: From
445: Eqs.~(\ref{eq:equil}, \ref{eq:single}) one can derive the following
446: equations linking
447: $u_m(\br)$, $\Pi_m(\br)$, and $\varepsilon_m$
448: (with $\alpha,\beta=1,2$):
449: \begin{subequations}
450: \label{eq:cross}
451: \begin{align}
452: \label{eq:cross_YL}
453: \nabla^2 {u}_m &= - \frac{2}{\gamma} {\Pi}_m ,
454: & &\br \in \hsmen , \\
455: \label{eq:cross_bc}
456: \bn_\alpha \cdot \nabla u_m
457: - \frac{u_m - \langle u_m \rangle_\alpha}{r_0} &=
458: \varepsilon_m
459: - \bn_\alpha \cdot \nabla u_\beta
460: + \frac{u_\beta - \langle u_\beta \rangle_\alpha}{r_0} ,
461: & &\br \in \partial S_\alpha
462: \quad (\beta \neq \alpha) , \\
463: {u}_m({\bf r}) &= 0 ,
464: & &\br \in C_L .
465: \end{align}
466: \end{subequations}
467: The superposition approximation is violated even if $\Pi_m=0$ and
468: $\varepsilon_m=0$ because of the boundary conditions at the contact
469: lines (Eq.~(\ref{eq:cross_bc})), as pointed out in \rcite{ODD05a}. The
470: case of non-vanishing $\Pi_m$ was addressed in
471: \rcites{ODD05b,WuFo05}.
472:
473: The capillary--induced effective interaction energy is defined as $\vmen (d) :=
474: \hat{\cal F}_{\rm eq} - 2 {\cal F}_{\rm eq}$, where $2 {\cal F}_{\rm
475: eq}$ is the sum of the equilibrium free energies of the
476: single--colloid
477: configurations, i.e., for $d\to\infty$.
478: $\vmen(d)$ depends parametrically on the (lateral) separation $d$
479: of the colloid centers in the reference configuration. This is {\em
480: not} the total interaction potential, which must include, e.g., the
481: direct electrostatic repulsion between charged colloids, not
482: considered in the expression~(\ref{eq:freeF}) for the free energy.
483: %
484: The effective interaction energy can be written as $\vmen = V_{\rm
485: sup} + V_{\rm corr}$, where $V_{\rm sup}$ is the result of imposing the
486: superposition approximation (Eq.~(\ref{eq:superpositionAnsatz})),
487: and $V_{\rm corr}$ is the
488: correction to this approximation. If we use the
489: expression~(\ref{eq:freeF1}), we obtain
490: \begin{subequations}
491: \label{eq:Vmen1}
492: \begin{equation}
493: \label{eq:Vsup1}
494: \avsup =
495: - \gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
496: (\nabla u_1) \cdot (\nabla u_2)
497: + \gamma \int_{S_1} \!\!\! dA \; |\nabla u_2|^2
498: - 2 \pi \gamma \left\langle u_2^2 - \langle u_2 \rangle_1^2
499: \right\rangle_1 ,
500: \end{equation}
501: \begin{align}
502: \label{eq:Vcorr1}
503: \avcorr =&
504: - \frac{1}{2} \gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
505: (\nabla u_m) \cdot \nabla (u_m + 4 u_2)
506: \nonumber \\
507: & - 2 \pi \gamma \left\langle u_m (u_m + 2 u_2)
508: - \langle u_m \rangle_1 \langle u_m + 2 u_2 \rangle_1
509: \right\rangle_1
510: - 2 \pi \gamma r_0^2 \varepsilon_m ( 2\epsf + \varepsilon_m) .
511: \end{align}
512: \end{subequations}
513: On the other hand, using expression~(\ref{eq:freeF2}), we
514: obtain
515: \begin{subequations}
516: \label{eq:Vmen2}
517: \begin{equation}
518: \label{eq:Vsup2}
519: \bvsup =
520: - \int_{\hsmen} \!\!\!\!\!\! dA \; \Pi_1 u_2
521: + \int_{S_1} \!\!\! dA \; \Pi_2 u_2
522: + 2 \pi \gamma r_0 \, \epsf \langle u_2 \rangle_1 ,
523: \end{equation}
524: \begin{align}
525: \label{eq:Vcorr2}
526: \bvcorr =&
527: - \int_{\hsmen} \!\!\!\!\!\! dA \;
528: \left[\Pi_2 \, u_m + 2 \Pi_m \, u_2
529: + \Pi_m \, u_m \right]
530: \nonumber \\
531: & + 2 \pi \gamma r_0 \langle (\epsf + \varepsilon_m) u_m
532: + \varepsilon_m (u_1 + u_2) \rangle_1
533: - 2 \pi \gamma r_0^2 \varepsilon_m ( 2\epsf + \varepsilon_m) .
534: \end{align}
535: \end{subequations}
536: We emphasize that the two alternative expressions of $V_{\rm sup}$ or
537: $V_{\rm corr}$ are not equivalent (but their sum $\vmen$ is), and they
538: in turn differ from $V_{\rm sup}$ as computed in \rcite{ODD05a} (see
539: Eq.~(39) therein), which was derived by inserting the superposition
540: ansatz directly into Eq.~(\ref{eq:freeF}) here. Application of Gauss'
541: theorem with Eq.~(\ref{eq:single_a}) leads to
542: \begin{equation}
543: \label{eq:discrepancy}
544: \avsup - \bvsup = 2\pi\gamma \left\langle r_0 (u_1 + u_2)
545: \frac{\partial u_2}{\partial n_1} -
546: (u_2 - \langle u_2 \rangle_1)^2 \right\rangle_1
547: = \bvcorr-\avcorr .
548: \end{equation}
549: The superposition approximation is inconsistent asymptotically in cases
550: in which $\avsup - \bvsup$ does not decay more rapidly than $\avsup$ as
551: function of the separation $d$. As remarked in \rcite{ODD05a}, there
552: are indeed relevant cases in which this consistency condition is not
553: fulfilled (see Subsec.~\ref{sec:asymptotic}).
554:
555: \subsection{Effective potential in the intermediate asymptotic regime $r_0 \ll d \ll L$}
556: \label{sec:asymptotic}
557:
558: In this subsection we compute $\vmen(d)$ asymptotically in the
559: intermediate range $r_0 \ll d \ll L$. For this purpose, we have to
560: make some restricting assumptions which, however, seem to be satisfied
561: in the experimental setups investigated so far.
562: First, in view of the discussion in the Introduction concerning the
563: electrical fields, we assume the proportionality
564: \begin{equation}
565: \label{eq:scalingPim}
566: \Pi_m^2 \sim |\Pi_1 \Pi_2| .
567: \end{equation}
568: This is valid if the interface stress is quadratic in a field
569: satisfying linear superposition in the two--particle configuration.
570: Examples are given below by some specific electrostatic models (see,
571: c.f.,
572: Eqs.~(\ref{eq:Pim_ideal}) and (\ref{eq:Pim_debye})).
573: Second, we assume that the single--colloid pressure $\Pi$ decays far
574: from the colloid as
575: \begin{equation}
576: \label{eq:Pidecay}
577: \Pi(r) \sim r^{-n} , \qquad n>4.
578: \end{equation}
579: In the experimentally relevant case of charged particles at water
580: interface, it has been established both theoretically \cite{Hurd85}
581: and experimentally \cite{ABCF02a} that $n=6$.
582: (This can be understood easily: the charge of a particle induces a
583: screening image charge in the water, so that the distant electric
584: field is dipolar.)
585: %
586: The constraint $n>4$ will allow us to estimate the
587: integrals appearing in Eqs.~(\ref{eq:Vmen1},~\ref{eq:Vmen2}) by
588: approximating them by the contribution of the regions near the
589: colloids. This condition excludes, however, the effect of an external
590: electric field (in that case the resulting pressure does not have to decay at all)
591: and the case that the colloidal charge is not perfectly screened so
592: that the distant electric field corresponds to a monopole, i.e., $n=4$
593: (this would occur if both fluid phases are dielectric, e.g., air and
594: insulating oil).
595:
596: The quantity $\hepsp - 2 \hepsf$ is the net (vertical) force by an
597: external agent acting on the total system consisting of two colloids
598: plus the interface (see Appendix~\ref{app:ufunctions}).
599: This can be, e.g., gravity (if the colloid is large enough for it to
600: be quantitatively relevant), dispersion forces by a substrate closely
601: beneath the interface (this effect can be modeled
602: similarly as gravity, see Appendix~\ref{app:flot}), or an optical tweezer
603: pushing the colloid vertically. Since we have assumed previously that
604: this external force is independent of the positions of the particles,
605: it is given by the sum of the net forces in the single--particle
606: configuration (i.e., if they are infinitely far apart from each
607: other):
608: \begin{equation}
609: \label{eq:additiveF}
610: \hepsp - 2\hepsf = 2 (\epsp - \epsf) .
611: \end{equation}
612: From the definitions in Eqs.~(\ref{eq:epsfp},~\ref{eq:hpi}) we can
613: write
614: \begin{subequations}
615: \begin{equation}
616: \label{eq:new_hepsp}
617: \hepsp = 2 (\epsp + \epsm - \varepsilon_{12})
618: \end{equation}
619: with
620: \begin{equation}
621: \label{eq:epsm_def}
622: \epsm := \frac{1}{2\pi \gamma r_0} \int_{\hsmen} \!\!\!\!\!\! dA \; \Pi_m ,
623: \qquad
624: \varepsilon_{12} := \frac{1}{2\pi \gamma r_0} \int_{{S}_1} \!\!\! dA \; \Pi_2 .
625: \end{equation}
626: \end{subequations}
627: For $d\to\infty$, it is clear that
628: \begin{equation}
629: \varepsilon_{12} \sim \frac{1}{d^{n}} .
630: \end{equation}
631: In this limit, we note that $\epsm$ receives its main contribution
632: from the regions around $S_\alpha$ implying
633: \begin{equation}
634: \label{eq:approx_epsm}
635: \epsm \sim \frac{2}{\gamma r_0}
636: \int^\infty_{r_0} \!\!\!\!\!\! dr \; r \, \Pi_m (r)
637: \sim \frac{1}{d^{n/2}}
638: \int^\infty_{r_0} \!\!\!\!\!\! dr \; \frac{r}{r^{n/2}} ,
639: \end{equation}
640: provided $n>4$, because $|\Pi_m (\br)| \sim \sqrt{\Pi(d) \, \Pi
641: (|{\bf r}-{\bf r}_\alpha|)}$ in those regions which provide the
642: dominant contribution to the integral.
643: %
644: Therefore, from Eqs.~(\ref{eq:hepsf}, \ref{eq:additiveF})
645: one obtains asymptotically
646: \begin{equation}
647: \label{eq:addF}
648: \varepsilon_m = \epsm - \varepsilon_{12} \sim \epsm \sim
649: \frac{1}{d^{n/2}} .
650: \end{equation}
651:
652: With these simplifying assumptions, we shall compute analytically the
653: behavior of $\vmen (d\to\infty)$ to leading order in $1/d$. More
654: precisely, on dimensional
655: grounds\footnote{Due to the rapid decay of $\hat{\Pi}$ far from the
656: particles, in the limit
657: $L\to\infty$ the length scale $L$ enters into the problem only if
658: $\varepsilon_\Pi \neq \varepsilon_F$ and in that case just as an
659: upper bound
660: on $d$ to avoid a logarithmic divergence (see the discussion after
661: Eq.~(\ref{eq:single_u})).} the expansion parameter is $r_0/d$.
662: In principle, $\hat{\Pi}$ can contain and thus introduce additional
663: length scales, for example the Debye length if the fluid phase is an
664: electrolyte. This complicates the problem, which then has to be
665: analyzed numerically
666: (see Subsec.~\ref{sec:debye}). \\
667:
668: \noindent {\bf The superposition approximation}: Within the
669: superposition approximation $\vmen(d)$ was computed in detail in
670: \rcite{ODD05a}. Here we sketch briefly the estimate of the asymptotic
671: behavior of $\bvsup$ (compare the three terms in
672: Eq.~(\ref{eq:Vsup2})):
673: \begin{subequations}
674: \begin{equation}
675: \int_{\hsmen} \!\!\!\!\!\! dA \; \Pi_1 u_2
676: \sim
677: 2\pi\gamma r_0 \epsp u(d) +
678: \Pi(d) \left[ 2 \int_{{S}_{\rm men, 2}} \!\!\!\!\!\! dA \; u_2
679: - \frac{1}{2} \pi r_0^3 \epsp + \pi r_0^2 \langle u_2 \rangle_2
680: \right] ,
681: \end{equation}
682: because the main contribution stems from the regions around
683: $S_\alpha$, and
684: \begin{equation}
685: \int_{S_1} \!\!\! dA \; \Pi_2 u_2
686: \sim
687: \Pi(d) u(d) \int_{S_1} \!\!\! dA \; ,
688: \end{equation}
689: \begin{equation}
690: 2\pi\gamma r_0 \epsf \langle u_2 \rangle_1
691: \sim 2\pi\gamma r_0 \epsf \left[ u(d) -
692: \frac{1}{4 \gamma} r_0^2 \, \Pi(d) \right] ,
693: \end{equation}
694: after expanding around $r_2 = d$. Thus
695: \begin{align}
696: \bvsup(d) &\sim 2\pi\gamma r_0 (\epsf - \epsp) u(d)
697: - \Pi(d) \left[
698: 2 \int_{{S}_{\rm men, 2}} \!\!\!\!\!\! dA \; u_2
699: + \pi r_0^2 \langle u_2 \rangle_2 \right] ,
700: \end{align}
701: \end{subequations}
702: because $\Pi(d)$ is asymptotically subdominant compared with $u(d)$
703: (see Eqs.~(\ref{eq:Pidecay}, \ref{eq:single_u})).
704: %
705: The difference $^a V_{\rm sup}-^b V_{\rm sup}$ in
706: Eq.~(\ref{eq:discrepancy}) between the two implementations of the
707: superposition approximation can be estimated as \cite{ODD05a} (see
708: Eq.~(\ref{eq:single_u}))
709: \begin{subequations}
710: \label{eq:discr_asymp}
711: \begin{align}
712: \left\langle (u_1 +u_2) \frac{\partial u_2}{\partial n_1}
713: \right\rangle_1
714: &\approx \frac{1}{2} r_0 \langle u_1 \rangle_1 \nabla^2 u(d)
715: \propto \Pi(d) \sim \frac{1}{d^n} , \\
716: \left\langle (u_2 - \langle u_2 \rangle_1)^2 \right\rangle_1
717: &\approx \frac{1}{2} r_0^2 |\nabla u(d)|^2
718: \sim \left\{
719: \begin{aligned}
720: 1/d^{2(n-1)} ,
721: & \quad \textrm{if } \epsf-\epsp = 0 , \\
722: 1/d^2 ,
723: & \quad \textrm{if } \epsf-\epsp \neq 0 .
724: \end{aligned}
725: \right.
726: \end{align}
727: \end{subequations}
728: Thus, whenever $\epsf-\epsp \neq 0$ (i.e., the system is not
729: mechanically isolated in the sense that there must be a force acting
730: on the boundary $C_L$ of the interface to compensate this non--vanishing net
731: force) one finds
732: \begin{equation}
733: \label{eq:vsup_asymp}
734: \bvsup(d) \sim 2\pi\gamma r_0 (\epsf - \epsp)
735: u(d) \sim 2\pi\gamma r_0^2 (\epsf-\epsp)^2 \ln\frac{d}{L} +
736: \rm{const} ,
737: \end{equation}
738: corresponding to an attractive force irrespective of the precise form
739: of the function $\Pi(r)$.
740: %
741: (The additive constant, which does not affect the physical
742: conclusions, depends on the precise form of the
743: boundary condition at $C_L$.)
744: %
745: Physically, $\bvsup$ is the work done by the net force
746: $2\pi\gamma(\epsf-\epsp)$ upon a vertical shift of the subsystem
747: consisting of one colloid plus its surrounding
748: interface (behaving like an ``effective particle'')
749: by an amount $u(d)$ due to the deformation induced by the
750: second colloid. In this case (i.e., $\epsf-\epsp \neq 0$),
751: the difference
752: $\avsup-\bvsup$
753: decays more rapidly than $\bvsup (d)$ and both expressions
754: $\avsup$ and $\bvsup$ agree asymptotically.
755: %
756: Equation~(\ref{eq:vsup_asymp}) exhibits the same dependence on $d$ as the
757: potential energy associated with the flotation force; this is
758: discussed briefly in Appendix~\ref{app:flot}.
759:
760: If $\epsf-\epsp=0$ (corresponding to mechanical isolation),
761: $\bvsup(d\to\infty) \sim 1/d^n$ (see Eq.~(\ref{eq:Pidecay})). This decay agrees with previous
762: findings \cite{MeAi03,FoWu04,ODD05a}, but the reliability of this
763: result is unclear because the difference $\avsup-\bvsup$
764: decays with the same power law. As a matter of fact, the amplitudes of
765: the asymptotic decay of $\avsup$ and $\bvsup$ differ and are in turn
766: different from Eq.~(52) in \rcite{ODD05a}, because there actually
767: neither of the two representations of ${\cal \hat{F}}_{\rm eq}$ was
768: used. However, for the leading behavior this is unimportant because
769: $\vmen$ is asymptotically dominated by the
770: correction to the superposition approximation. \\
771:
772: \noindent {\bf Beyond the superposition approximation}: If the
773: interface deformation field in Eq.~(\ref{eq:hu}) is evaluated near
774: colloid 1, the term $u_2+u_m$ is dominated by $u_2$ in the absence of
775: mechanical isolation and by $u_m$ in the case of mechanical isolation
776: (see, c.f., Eqs.~(\ref{eq:single_u}, \ref{eq:cross_u})):
777: \begin{equation}
778: u_2 + u_m \sim \left\{
779: \begin{aligned}
780: u_m(d) \sim d^{-n/2} ,
781: & \quad \textrm{if } \epsf-\epsp = 0 , \\
782: u_2(d) \sim \ln d ,
783: & \quad \textrm{if } \epsf-\epsp \neq 0 .
784: \end{aligned}
785: \right.
786: \end{equation}
787: Thus the superposition approximation holds if and only if
788: the system is {\em
789: not} mechanically isolated. Otherwise, the correction $u_m$ is
790: dominant and, for asymptotically large separations $d$, $\vmen \sim
791: V_{\rm corr}$. Hence in the following we take $\epsf-\epsp=0$. Since
792: the main contributions to the integrals stem from the regions around
793: $S_\alpha$ one obtains the estimates
794: \begin{equation}
795: - \int_{\hsmen} \!\!\!\!\!\! dA \;
796: \left[\Pi_2 \, u_m + 2 \Pi_m \, u_2 + \Pi_m \, u_m \right]
797: \sim - \int_{{S}_{\rm men, 2}} \!\!\!\!\!\! dA \;
798: \left[\Pi_2 \, u_m + 2 \Pi_m \, u_2 \right] \sim \frac{1}{d^{n/2}} .
799: \end{equation}
800: Using the asymptotic decay of $u_m$ given by Eq.~(\ref{eq:cross_u}) we
801: find
802: \begin{equation}
803: \label{eq:bvcorr_asymp}
804: \bvcorr(d) \sim
805: - \int_{{S}_{\rm men, 2}} \!\!\!\!\!\! dA \;
806: \left[\Pi_2 \, u_m + 2 \Pi_m \, u_2 \right]
807: + 2 \pi \gamma r_0 \left\langle \epsf u_m + \epsm u_1
808: \right\rangle_1
809: - 4 \pi \gamma r_0^2 \epsf \epsm
810: \sim \frac{1}{d^{n/2}} ,
811: \end{equation}
812: because each term scales like $d^{-n/2}$ and there is no reason for
813: mutual cancellations. In this expression one can identify two distinct
814: contributions with a simple physical meaning. After inserting
815: Eqs.~(\ref{eq:hu},~\ref{eq:hepsf}) into the
816: relationship~(\ref{eq:equil_h}) we write $\Delta\hat{h}_2 = \Delta h_2
817: + \Delta h_m$ with
818: \begin{subequations}
819: \begin{align}
820: \Delta h_2 &:= \langle u_1 + u_2 \rangle_2 - r_0 \, \epsf ,
821: \quad (\sim \langle u_2 \rangle_2 - r_0 \, \epsf + {\cal O}(d^{-n})) \\
822: \Delta h_m &:= \langle u_m \rangle_2 - r_0 \, \varepsilon_m
823: \quad (\sim d^{-n/2}) ,
824: \end{align}
825: \end{subequations}
826: where the asymptotic decays are giving by Eqs.~(\ref{eq:addF},~\ref{eq:cross_u}).
827: Accordingly, expression~(\ref{eq:bvcorr_asymp}) shows that $\bvcorr$ is
828: dominated asymptotically by\footnote{Note that by symmetry $\langle
829: u_2 \rangle_2 = \langle u_1 \rangle_1$ and $\langle u_m \rangle_2 =
830: \langle u_m \rangle_1$.} (i) the work done by the additional pressure
831: $2\Pi_m$ if the single--colloid configuration is deformed relative to
832: the reference configuration, that is, $-\int dA \; 2 \Pi_m \,
833: u_2 + 2\pi \gamma r_0 \epsm \Delta h_2$,
834: and (ii) the work done by the
835: forces acting in the single--colloid configuration upon the additional displacement $u_m$, that is, $-\int dA
836: \; \Pi_2 \, u_m + 2\pi \gamma r_0 \epsf \Delta h_m$.
837:
838:
839: The asymptotic behavior of $\avcorr(d\to\infty)$ can be derived by
840: estimating the behavior of the terms in Eq.~(\ref{eq:Vcorr1})
841: individually, as in the case of Eq.~(\ref{eq:Vcorr2}), with the result
842: \begin{equation}
843: \label{eq:avcorr_asymp}
844: \avcorr(d) \sim
845: - 2 \gamma \int_{S_{\rm men, 2}} \!\!\!\!\!\! dA \;
846: (\nabla u_m) \cdot (\nabla u_2)
847: - 4 \pi \gamma r_0^2 \epsf \epsm
848: \sim \frac{1}{d^{n/2}} .
849: \end{equation}
850: The asymptotic decay predicted by Eqs.~(\ref{eq:bvcorr_asymp}) and
851: (\ref{eq:avcorr_asymp}) must agree both with respect to the decay
852: exponents and the amplitudes, because the difference $^a V_{\rm
853: sup}-^b V_{\rm sup}$ decays asymptotically more rapidly (see
854: Eq.~(\ref{eq:discr_asymp}) for $\epsf-\epsp = 0$).
855: %
856: The sign of the force described by $V_{\rm corr}(d)$ is not evident
857: from the outset, but in the applications we shall consider later, it
858: turns out to be always attractive.
859:
860: It is interesting to compare our result with the corresponding one in
861: \rcite{WuFo05}. After noting the equivalence $u_m \leftrightarrow 2
862: u_{12}$ in the notations, one finds that Eq.~(7) in \rcite{WuFo05} is
863: identical with the integral term in Eq.~(\ref{eq:avcorr_asymp}) here.
864: We obtain an additional term $\propto \epsf \epsm$ because we treat
865: the contribution by the colloid to the free energy functional in full
866: detail, while in \rcite{WuFo05} the point--particle approximation
867: ($r_0 \rightarrow 0$) is used from the outset (compare
868: Eq.~(\ref{eq:freeF1}) here with Eq.~(1) in \rcite{WuFo05}).
869: %
870: As a consequence, in \rcite{WuFo05} our additional term is lost as a
871: singularity of the integral term, which is regularized there by introducing
872: an unknown cutoff length expected to be of the order of $r_0$ (see
873: also \rcites{FoWu04,ODD05a}).
874: %
875: The analysis of realistic models in Sec.~\ref{sec:appl} will show that
876: the quantitative contribution of the term $\propto \epsf \epsm$ to the
877: effective interaction $\vmen$ is actually larger than but
878: proportional to the contribution from the other term in
879: Eq.~(\ref{eq:avcorr_asymp}).
880: \\
881:
882: To summarize, the capillary--induced effective interaction between two
883: colloids is given by (compare Eq.~(\ref{eq:Pidecay}) defining $n$)
884: \begin{equation}
885: \label{eq:vmen_asymp}
886: \vmen (d) \sim \left\{
887: \begin{aligned}
888: V_{\rm corr}(d) & \sim d^{-n/2} , \\
889: V_{\rm sup}(d) & \sim \ln d ,
890: \end{aligned}\qquad
891: \begin{aligned}
892: & \textrm{if } \epsf-\epsp = 0 , \\
893: & \textrm{if } \epsf-\epsp \neq 0 ,
894: \end{aligned}
895: \right. \qquad (r_0 \ll d) .
896: \end{equation}
897:
898:
899:
900: \section{Applications}
901: \label{sec:appl}
902:
903: In this section we compute $\vmen(d)$ for $\epsf=\epsp$ (mechanical
904: isolation) for different realistic models of $\hat{\Pi}$ derived from
905: the solution of the electrostatic problem within various
906: approximations. We note that the asymptotic decay of $\vmen(d)$ is the
907: same as that of the direct electrostatic repulsion, so that these
908: detailed calculations beyond the asymptotic analysis of
909: Subsec.~\ref{sec:asymptotic} are necessary in order to be able to
910: address this
911: fine--tuning problem and to determine whether the total force is
912: asymptotically attractive or repulsive.
913:
914: \subsection{Ideally conducting fluid phase}
915: \label{sec:idealconductor}
916:
917: The simplest model consists of approximating water by an ideal
918: conductor. Formally this corresponds to the limit of zero
919: temperature, so that the osmotic pressure of the mobile charges
920: accumulated at the interface vanishes and the Debye length
921: $\kappa^{-1}$ is zero (see, c.f., Eq.~(\ref{eq:debyelength})). In this case the electric field $\hat{\bf E} = \hat{E}
922: \ez$ is always normal to the interface and the pressure is given by
923: Maxwell's stress tensor evaluated at the insulating side of the
924: interface (we use Gaussian units),
925: \begin{equation}
926: \hat{\Pi} = \frac{1}{4\pi} \left. \ez \cdot \left(
927: \epsilon \hat{\bf E} \hat{\bf E} -
928: \frac{1}{2} \epsilon \hat{E}^2 {\bf 1}
929: \right) \cdot \ez \right|_{z=0^+} =
930: \frac{\epsilon_1}{8\pi} \hat{E}^2 (z=0^+) ,
931: \end{equation}
932: where $\epsilon_1$ is the dielectric constant of the insulating phase
933: ($z>0$).
934:
935: In the present context the electrostatic problem of a charged sphere partially immersed in a
936: conducting fluid has been solved numerically and
937: semi--analytically in \rcite{DaKr06a}. There it has been shown that
938: the single--colloid pressure field exhibits an integrable divergence
939: upon approaching the three--phase contact line and that asymptotically
940: it displays the familiar dipole behavior. The following approximate
941: parametrization (Eq.~(1.4) in \rcite{DaKr06a} expressed in terms of
942: our notation) incorporates these properties and is sufficiently
943: accurate for our present purposes:
944: \begin{equation}
945: \label{eq:fitPi}
946: \Pi(r) = \frac{\gamma \epsf}{r_0} \, b(\mu)
947: \left[ \frac{r}{r_0}-1 \right]^{\mu-1}
948: \left[ \frac{r}{r_0} \right]^{-\mu-5} , \qquad
949: b(\mu) := \frac{1}{6}\mu(\mu+1)(\mu+2)(\mu+3) ,
950: \end{equation}
951: where $0<\mu<1$ is a fitting parameter the precise value of which
952: depends on
953: the contact angle $\theta$ and the dielectric constant $\epsilon_1$.
954: This expression is normalized so that $\epsp=\epsf$ and it corresponds
955: to an exponent $n=6$ independent of the choice for $\mu$.
956: Therefore, far from a colloid, the
957: single--colloid electric field $|{\bf E}|=\sqrt{8\pi\Pi/\epsilon_1}$ is that
958: of a dipole perpendicular to the interface, the strength of which is
959: given by
960: \begin{equation}
961: \label{eq:dipole}
962: |p| = \lim_{r \to \infty} \epsilon_1 \, r^3 \,
963: |{\bf E}(r)| =
964: \epsilon_1 \, r_0^3 \,
965: \sqrt{\frac{8\pi\gamma\epsf b(\mu)}{\epsilon_1 r_0}} .
966: \end{equation}
967:
968: In the presence of two colloids, we take $\hat{{\bf E}} \approx {\bf
969: E}_1 + {\bf E}_2$. This approximation allows a simplification of the
970: calculations and should not alter the physical picture significantly.
971: The approximative character is due to possible violations of the
972: electrostatic boundary conditions at the surfaces of the colloidal particles: The
973: additional polarization of colloid 1 induced by ${\bf E}_2$ will
974: actually lead to an electric field $\hat{{\bf E}} = {\bf E}_1 + {\bf
975: E}_2 + \delta {\bf E}$ in the neighborhood of colloid 1, with an
976: induced electric field $|\delta {\bf E}(\br,d)| \sim \chi(\br) E_2
977: (d)$, where the ``effective susceptibility'' $\chi(\br)$ is expected to be at most of
978: order unity. Thus, our conclusions will be qualitatively correct with
979: a quantitative error of a factor of order unity.
980: %
981: Under these conditions the field $\Pi_m(\br)$ defined by Eq.~(\ref{eq:hpi}) is given
982: within this approximation by
983: \begin{equation}
984: \label{eq:Pim_ideal}
985: \Pi_m (\br) = \sqrt{\Pi (|\br - \br_1|) \Pi (|\br - \br_2|)} .
986: \end{equation}
987: The integrals in Eq.~(\ref{eq:avcorr_asymp}) are computed by using the
988: expressions in Eqs.~(\ref{eq:single_u}, \ref{eq:cross_u}) so that
989: in the limit $r_0 \ll d$ one obtains (for details see
990: Appendix~\ref{app:ideal_calc})
991: \begin{subequations}
992: \label{eq:ideal_int}
993: \begin{equation}
994: \epsm \approx \frac{4 b(\mu)}{\mu+1} \epsf
995: \left(\frac{r_0}{d}\right)^3
996: \end{equation}
997: and
998: \begin{equation}
999: \int_{S_{\rm men, 2}} \!\!\!\!\!\! dA \;
1000: (\nabla u_m) \cdot (\nabla u_2) \approx
1001: \frac{8\pi b(\mu)}{\mu+1} M(\mu) \, r_0^2 \epsf^2
1002: \left(\frac{r_0}{d}\right)^3 ,
1003: \end{equation}
1004: \end{subequations}
1005: with
1006: \begin{equation}
1007: \label{eq:Mfunction}
1008: M(\mu) := \frac{1}{8}(\mu+1) b(\mu)
1009: \int_0^1 \!\!\! dv \; v^4
1010: \, _2F_1(1-\mu,4;5;v) \,
1011: _2F_1(\frac{1-\mu}{2},1;2;v)
1012: \end{equation}
1013: and $_2F_1(\alpha,\beta;\gamma;z)$ is the
1014: hypergeometric function (see Subsec.~9.1 in \rcite{GrRy94}).
1015: %
1016: We have checked numerically that $M(\mu) \approx \mu/5$ with a maximum
1017: error of $\approx 1.3\%$ within the range $0\leq\mu\leq 1$. The final
1018: result reads
1019: \begin{equation}
1020: \label{eq:approxVmen}
1021: \vmen (d) \approx \avcorr(d) \approx
1022: - \frac{16\pi b(\mu)}{\mu+1} [1 + M(\mu)] \,
1023: \gamma r_0^2 \, \epsf^2
1024: \left(\frac{r_0}{d}\right)^3 .
1025: \end{equation}
1026: This corresponds to an attractive force. We note that the dominant
1027: contribution to $\vmen (d)$ stems from the term proportional to $\epsm$ in
1028: Eq.~(\ref{eq:avcorr_asymp}), the more so the smaller $\mu$ is.
1029: Smaller values of $\mu$ correspond to an increasing importance of the
1030: electric field near the colloid (see Eq.~(\ref{eq:fitPi})).
1031:
1032: $\vmen(d)$ is to be compared with the potential energy due to the
1033: direct electrostatic repulsion $V_{\rm rep}(d)$ of the charged
1034: colloids. The potential energy of one dipole in the field of another
1035: identical dipole is $p^2 /(\epsilon_1 d^3)$ so that $V_{\rm
1036: rep}(d) \approx p^2/(2 \epsilon_1 d^3)$ for large $d$. (One must
1037: divide by a factor $2$ because within our model no work is done on the
1038: image charge inside the conducting phase forming the dipole.)
1039: Collecting the results, we find that the total interaction energy at
1040: large separations is given by
1041: \begin{equation}
1042: \label{eq:Vtotal}
1043: V_{\rm total}(d) = V_{\rm rep}(d) + \vmen(d) \approx
1044: 4 \pi \gamma r_0^2 \, \epsf b(\mu)
1045: \left(\frac{r_0}{d}\right)^3 \left[ 1
1046: - \frac{4}{\mu+1} [1+M(\mu)] \epsf \right] .
1047: \end{equation}
1048: Hence we see that the attractive capillary potential is proportional to
1049: $\epsf^2$ and the direct electrostatic repulsion is proportional to $\epsf$.
1050: Thus in the limit $\epsf \ll 1$, on which our calculations are
1051: based, the electrostatic repulsion is always larger than the capillary
1052: attraction. The leading--order analysis of this model predicts an
1053: attraction only if the charge of the colloid is large enough so that
1054: $\epsf = {\cal O}(1)$. The critical value above which there is
1055: attraction is given by
1056: \begin{equation}
1057: \label{eq:critical}
1058: \varepsilon_{F, \rm crit} (\mu)= \frac{\mu+1}{4[1+M(\mu)]} ,
1059: \end{equation}
1060: which lies in the range $1/4 < \varepsilon_{F, \rm crit} < 5/12$ for
1061: $0<\mu<1$.
1062:
1063:
1064: \subsection{Finite Debye length}
1065: \label{sec:debye}
1066:
1067: Consider now the more general case of an upper insulating phase
1068: (dielectric constant $\epsilon_1$) and a lower electrolytic phase
1069: (dielectric constant $\epsilon_2$, electrolyte concentration $n_0$) at
1070: a finite temperature $T$. The stress tensor acting on the interface is
1071: due to the difference of Maxwell's stress tensor just above and below
1072: the interface, plus an osmotic pressure $\hat{p}_{\rm osm}$ by the
1073: excess of ions concentrated close to the interface:
1074: \begin{equation}
1075: \hat{\Pi} = \frac{1}{4\pi} \left. \ez \cdot \left(
1076: \epsilon \hat{\bf E} \hat{\bf E} -
1077: \frac{1}{2} \epsilon \hat{E}^2 {\bf 1}
1078: \right) \cdot \ez \right|^{z=0^+}_{z=0^-}
1079: + \hat{p}_{\rm osm} .
1080: \end{equation}
1081: With $\hat{\Phi}(\br)$ denoting the electrostatic potential at the
1082: interface, $\hat{E}_{z,\pm}(\br) := \ez \cdot \hat{\bf
1083: E}(\br,z=0^\pm)$ as the normal component of the electric field at the
1084: interface (with $\epsilon_1 \hat{E}_{z,+}=\epsilon_2 \hat{E}_{z,-}$)
1085: and $\Ep(\br, z=0)$ as the (continuous) parallel component at the
1086: interface we have
1087: \begin{equation}
1088: \label{eq:maxwellPi}
1089: \frac{1}{4\pi} \left. \ez \cdot \left(
1090: \epsilon \hat{\bf E} \hat{\bf E} -
1091: \frac{1}{2} \epsilon \hat{E}^2 {\bf 1}
1092: \right) \cdot \ez \right|^{z=0^+}_{z=0^-} =
1093: \frac{\epsilon_2-\epsilon_1}{8\pi} \left[
1094: \frac{\epsilon_1}{\epsilon_2} \Ezp^2 + \Ep^2
1095: \right],
1096: \end{equation}
1097: and with $\Delta n(\br)$ as the excess ion concentration at the interface
1098: \begin{equation}
1099: \label{eq:posm}
1100: \hat{p}_{\rm osm} \approx k_B T \Delta n =
1101: n_0 k_B T \left[ \exp\left(\frac{e\hat{\Phi}}{k_B T}\right) +
1102: \exp\left(- \frac{e\hat{\Phi}}{k_B T}\right) -2 \right] ,
1103: \end{equation}
1104: assuming that the electrolyte is dilute and consists of monovalent
1105: ions.
1106:
1107: In order to solve the electrostatic problem we introduce two
1108: simplifications: We apply the Debye--H\"uckel approximation for dilute
1109: electrolytes and we approximate the extended colloid by a point charge
1110: $q$ at its center, i.e., we retain only the monopole term of the
1111: source of the field. This 'monopolar' approximation corresponds
1112: formally to the limit of vanishing contact radius, $r_0 \rightarrow
1113: 0$,
1114: and it can be expected to provide the correct field at distances from
1115: the colloid which are large compared to $r_0$. This is in fact
1116: complementary to the limiting case $\kappa r_0 \gg 1$ considered in
1117: the previous subsection.
1118: %
1119: From Eq.~(\ref{eq:posm}) we find
1120: \begin{equation}
1121: \hat{p}_{\rm osm} \approx \frac{\epsilon_2}{8\pi}
1122: \kappa^2\,\hat{\Phi}^2 ,
1123: \end{equation}
1124: where the screening length is given by
1125: \begin{equation}
1126: \label{eq:debyelength}
1127: \kappa^{-1}=\sqrt{\frac{\epsilon_2 k_B T}{8\pi n_0 e^2}} .
1128: \end{equation}
1129: For pure water $\epsilon_2 = 81$ and $n_0 \approx 10^{-7}$ mol/l
1130: (that is, $\rm{pH}=7$)
1131: lead to a screening length $\approx 1\, \mu$m at room temperature.
1132:
1133: We consider first the case of a single colloid corresponding to a
1134: point charge $q$ located at the flat interface between an insulator
1135: and an electrolyte. Proceeding along the lines of, e.g., \rcites{Hurd85,ACNP00}, we find:
1136: \begin{subequations}
1137: \label{eq:Efields}
1138: \begin{eqnarray}
1139: \Phi (r) &=& \frac{2q}{\epsilon_2 r}\; {\cal I}_a(\kappa r) , \\
1140: E_{z,+}(r) &=& -\frac{2q}{\epsilon_2 r^2}\;
1141: {\cal I}_b(\kappa r) , \\
1142: {\bf E}_{\parallel} (r) = \mbox{} - \frac{d \Phi}{d r} \er &=&
1143: \frac{2q}{\epsilon_2 r^2}\;
1144: {\cal I}_c(\kappa r) \; \er,
1145: \end{eqnarray}
1146: \end{subequations}
1147: where $\er$ is the unit radial vector pointing away from the center of the
1148: colloid, and the auxiliary functions ${\cal I}_n(k)$ are given by
1149: integrals over a Bessel function:
1150: \begin{subequations}
1151: \label{eq:Idef}
1152: \begin{eqnarray}
1153: {\cal I}_a(k) &:=& \int_0^\infty dx\; J_0(x)\;
1154: \frac{x}{\frac{\epsilon_1}{\epsilon_2} x +
1155: \sqrt{x^2+k^2}} , \\
1156: {\cal I}_b(k) &:=& \frac{\epsilon_2}{\epsilon_1} \int_0^\infty dx\; J_0(x)\;
1157: \frac{x\,\sqrt{x^2+k^2}}{
1158: \frac{\epsilon_1}{\epsilon_2} x + \sqrt{x^2+k^2}} , \\
1159: {\cal I}_c(k) &:=& {\cal I}_a(k) - k\, \frac{d\, {\cal I}_a}{d k} (k).
1160: \end{eqnarray}
1161: \end{subequations}
1162: These integrals have been computed numerically; the analytic
1163: expressions for their asymptotic behaviors are derived in
1164: Appendix~\ref{app:Ifunctions}. Thus, taking $\epsilon_2/\epsilon_1 \gg 1$ (the
1165: ratio $\epsilon_2/\epsilon_1$ is approximately $81$ for
1166: water in contact with air)
1167: in the intermediate asymptotic regime $1 \ll \kappa r \ll
1168: (\epsilon_2/\epsilon_1)^2$ one obtains the following expressions for the potential
1169: and the field components (see Eq.~(\ref{eq:Efields})):
1170: \begin{subequations}
1171: \label{eq:asymp}
1172: \begin{eqnarray}
1173: \Phi(r) & \approx & \frac{2q}{\epsilon_2 r} \left[
1174: \frac{\epsilon_1}{\epsilon_2} \frac{1}{(\kappa r)^2}
1175: + {\rm e}^{-\kappa r} \right] , \\
1176: E_{z,+}(r) & \approx & \frac{2q}{\epsilon_2 r^2} \left[
1177: - \frac{1}{\kappa r} + \sqrt{\frac{\pi}{2}} \;
1178: \frac{\epsilon_1}{\epsilon_2} \; (\kappa r)^{3/2}\,
1179: {\rm e}^{-\kappa r} \right] , \\
1180: {\bf E}_{\parallel}(r) & \approx & \frac{2q}{\epsilon_2 r^2} \left[
1181: 3 \frac{\epsilon_1}{\epsilon_2} \frac{1}{(\kappa r)^2}
1182: + \kappa r \, {\rm e}^{-\kappa r} \right] \er .
1183: \end{eqnarray}
1184: \end{subequations}
1185: Asymptotically $\Phi(r)$ decays as $1/r^3$ and the electrostatic
1186: interaction energy $q \Phi (d)$ of a second charge $q$ at a distance
1187: $d$ from the first one decays likewise as $1/d^3$. This is the
1188: celebrated dipole repulsion between charged colloidal particles first
1189: conjectured in \rcite{Pier80}. However, due to the large value of the
1190: ratio $\epsilon_2/\epsilon_1$, closer to the charge there is a crossover to a screened
1191: Coulomb potential $\Phi(r) \propto \exp{(-\kappa r)}/r$. One can introduce a crossover length $r_{\rm cross}$
1192: defined from the asymptotic behavior given by Eq.~(\ref{eq:asymp}) as
1193: \begin{equation}
1194: \label{eq:rcross}
1195: \frac{\epsilon_1}{\epsilon_2} \frac{1}{(\kappa r_{\rm cross})^2}
1196: = {\rm e}^{-\kappa r_{\rm cross}} .
1197: \end{equation}
1198: This equation has solutions only if $\epsilon_1/\epsilon_2 < (2/{\rm
1199: e})^2 \approx 0.54$, in which case the relevant solution $\kappa
1200: r_{\rm cross}$ is larger than $2$ and depends only weakly, i.e., logarithmically on the precise value of the ratio
1201: $\epsilon_1/\epsilon_2$.
1202: For $\epsilon_2/\epsilon_1=81$ this gives
1203: $\kappa r_{\rm cross} \approx 8.7$. The parallel component
1204: ${\bf E}_{\parallel}$ exhibits the same crossover, but the normal component
1205: $E_{z,+}$ always decays algebraically and is much larger than the
1206: parallel component.
1207: %
1208: Nevertheless, the contribution of $E_{z,+}$ to the pressure field
1209: $\Pi(r)$ is reduced by a factor $\epsilon_1/\epsilon_2$ (see
1210: Eq.~(\ref{eq:maxwellPi})), so that $\Pi(r)$ will also exhibit the
1211: crossover at a distance $r \approx r_{\rm cross}$. From Eqs.~(\ref{eq:maxwellPi}, \ref{eq:posm}) one finds
1212: \begin{equation}
1213: \label{eq:singlepi}
1214: \Pi(r) = \gamma r_0 \kappa^2
1215: \frac{\epsf}{(\kappa r)^4 {\cal P} (\kappa r_0)} \left[
1216: \frac{\epsilon_1}{\epsilon_2}
1217: \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1218: {\cal I}_b^2(\kappa r)
1219: + \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1220: {\cal I}_c^2(\kappa r)
1221: + (\kappa r)^2 {\cal I}_a^2(\kappa r)
1222: \right] ,
1223: \end{equation}
1224: where (see Eq.~(\ref{eq:epsfp}))
1225: \begin{equation}
1226: \label{eq:qeff}
1227: \epsf = \epsp = \frac{q^2 \kappa^2}{2 \pi \epsilon_2 \gamma r_0}
1228: {\cal P}(\kappa r_0)
1229: \end{equation}
1230: with the dimensionless function
1231: \begin{eqnarray}
1232: \label{eq:calP}
1233: {\cal P}(\kappa r_0) & := &
1234: \int_{\kappa r_0}^{\infty} \!\! dx \;
1235: \frac{1}{x^3} \left[
1236: \frac{\epsilon_1}{\epsilon_2}
1237: \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1238: {\cal I}_b^2(x)
1239: + \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1240: {\cal I}_c^2(x) + x^2 {\cal I}_a^2(x)
1241: \right] \\
1242: & & \nonumber \\
1243: & \approx &
1244: \left[\frac{1}{\kappa r_0} + \frac{1}{2 (\kappa r_0)^2} \right]
1245: {\rm e}^{-2\,\kappa r_0} , \nonumber
1246: \end{eqnarray}
1247: where the last line is the leading (zeroth order) contribution in an expansion in terms of the small parameter
1248: $\epsilon_1/\epsilon_2$.
1249: %
1250: Equation~(\ref{eq:qeff}) allows one to express the
1251: parameter $q$, the value of which is often uncertain,
1252: in terms of the more convenient parameter $\epsf$.
1253: %
1254: Figure~\ref{fig:cross} shows how $\Pi(r)$ attains the dipole limiting
1255: behavior $\propto 1/r^6$ due to the electrostatic pressure $\propto
1256: E_{z,+}^2$ beyond the crossover distance; at shorter distances,
1257: $\Pi(r)$ is dominated by the term $\propto {\bf E}_{\parallel}^2$ and,
1258: to a lesser extent, by the osmotic pressure.
1259: %
1260: The figure indicates also that the main contribution to the total
1261: force $\epsp$ stems from the regions close to the particle, as
1262: evidenced also by the formal divergence of ${\cal P}(\kappa r_0 \to 0)$
1263: (see Eq.~(\ref{eq:calP})). Thus, the precise value of $\epsp$ will be
1264: affected by the fact that
1265: our solution of the electrostatic problem within the monopolar
1266: approximation is expected to be reliable in principle only sufficiently
1267: far from the particle.
1268:
1269: \begin{figure}
1270: \begin{center}
1271: \epsfig{file=figures/crossover_pi.eps,width=.5\textwidth}
1272: \caption{
1273: The dimensionless stress $\bar\Pi :=
1274: \Pi/[\gamma r_0 \kappa^2 \epsf/{\cal P}(\kappa r_0)]$
1275: due to a single point charge at the interface for $\epsilon_2/\epsilon_1=81$, as given by
1276: Eq.~(\ref{eq:singlepi}) (thick solid line). The thin lines
1277: represent each of the three additive contributions in
1278: Eq.~(\ref{eq:singlepi}): due to $E_{z}$, only given by ${\cal I}_b$;
1279: due to $E_{\parallel}$, only given by ${\cal I}_c$; and
1280: due to the osmotic pressure, only given by ${\cal I}_a$.
1281: As expected, one observes a crossover in $\bar\Pi$
1282: at a distance $r$ comparable with $r_{\rm cross}$ defined by
1283: Eq.~(\ref{eq:rcross}).}
1284: \label{fig:cross}
1285: \end{center}
1286: \end{figure}
1287:
1288: We consider now two identical point charges at the interface separated
1289: by a distance $d$. Within the Debye--H\"uckel and point--charge
1290: approximations, the solution of this electrostatic problem is given by
1291: the superposition of the single--colloid fields.
1292: In this case the field $\Pi_m(\br)$ in Eq.~(\ref{eq:hpi}) reads (as in
1293: Subsec.~\ref{sec:exact}, the subscript $\alpha(=1,2)$ denotes that the
1294: corresponding field is evaluated with respect to
1295: particle $\alpha$;
1296: see Fig.~\ref{fig:2coll_top} for the notation):
1297: \begin{eqnarray}
1298: \label{eq:Pim_debye}
1299: \Pi_m(r) & = &
1300: \frac{\epsilon_2 - \epsilon_1}{8\pi} \left[
1301: \frac{\epsilon_1}{\epsilon_2} (E_{z,+})_1 (E_{z,+})_2
1302: + ({\bf E}_{\parallel})_1 \cdot ({\bf E}_{\parallel})_2 \right]
1303: + \frac{\epsilon_2}{8\pi} \kappa^2 \Phi_1 \Phi_2 \nonumber \\
1304: & & \nonumber \\
1305: & = &
1306: \frac{q^2 \kappa^4}{2 \pi \epsilon_2}
1307: \frac{1}{\kappa^4 (|\br -\br_1| |\br -\br_2|)^2} \left[
1308: \frac{\epsilon_1}{\epsilon_2}
1309: \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1310: {\cal I}_b(\kappa |\br -\br_1|) {\cal I}_b(\kappa |\br -\br_2|)
1311: + \mbox{} \right. \nonumber \\
1312: & & \nonumber \\
1313: & & \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1314: {\cal I}_c(\kappa |\br -\br_1|) {\cal I}_c(\kappa |\br -\br_2|)
1315: \frac{\br -\br_1}{|\br -\br_1|} \cdot
1316: \frac{\br -\br_2}{|\br -\br_2|} + \mbox{} \nonumber \\
1317: & & \nonumber \\
1318: & & \left. \frac{}{} \kappa^2 |\br -\br_1| |\br -\br_2| \,
1319: {\cal I}_a(\kappa |\br -\br_1|) {\cal I}_a(\kappa |\br -\br_2|)
1320: \right] .
1321: \end{eqnarray}
1322: Therefore one has (see Eqs.~(\ref{eq:epsm_def}, \ref{eq:qeff}))
1323: \begin{equation}
1324: \epsm =
1325: \frac{{\cal P}_m(\kappa r_0, \kappa d)}{{\cal P}(\kappa r_0)}
1326: \, \epsf ,
1327: \end{equation}
1328: where
1329: \begin{eqnarray}
1330: \label{eq:defPm}
1331: {\cal P}_m (\kappa r_0, \kappa d)& := & \frac{1}{2\pi}
1332: \int_{\hat{S}_{\rm men}} \!\!\!\!\!\!\!\! d^2 {\bf x} \,
1333: \frac{1}{(x_1 x_2)^2}
1334: \left[ \frac{\epsilon_1}{\epsilon_2}
1335: \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1336: {\cal I}_b(x_1) {\cal I}_b(x_2) + \mbox{} \right. \\
1337: & & \nonumber \\
1338: & & \left. \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1339: {\cal I}_c(x_1) {\cal I}_c(x_2)
1340: \, \frac{{\bf x}_1 \cdot {\bf x}_2}{x_1 x_2}
1341: + x_1 x_2 \,
1342: {\cal I}_a(x_1) {\cal I}_a(x_2)
1343: \right] \nonumber
1344: \end{eqnarray}
1345: in terms of ${\bf x}:=\kappa \br$ and ${\bf x}_\alpha := \kappa (\br -
1346: \br_\alpha)$. Dependences on $r_0$ and $d$ enter through the
1347: specification of the integration domain $\hat{S}_{\rm men}$.
1348: %
1349: Using this expression and Eq.~(\ref{eq:Idef}), we have integrated
1350: numerically Eq.~(\ref{eq:cross}) using bipolar coordinates in a domain
1351: corresponding to a size $L \approx 100 / \kappa$. Taking this
1352: numerical solution and the exact expression in
1353: Eq.~(\ref{eq:single_u}), we have subsequently computed numerically the
1354: effective interaction energy as $\vmen(d)\approx \bvcorr(d)$ by using
1355: Eq.~(\ref{eq:Vcorr2}). As in the case studied in
1356: Subsec.~\ref{sec:idealconductor} (see the text after
1357: Eq.~(\ref{eq:approxVmen})), the last term in Eq.~(\ref{eq:Vcorr2})
1358: turns out to be larger in magnitude than the remaining terms but
1359: they are approximately proportional to each other, so that for the
1360: purpose of understanding the numerical results we can consider the
1361: following proportionality (upon the application of Eq.~(\ref{eq:addF}))
1362: \begin{equation}
1363: \label{eq:vmen_debye}
1364: \vmen(d) \propto
1365: \mbox{}- 2 \pi \gamma r_0^2 \varepsilon_m ( 2\epsf + \varepsilon_m)
1366: \sim - 4\pi \gamma r_0^2 \epsf \epsm =
1367: - 4\pi \gamma r_0^2 \epsf^2 \,
1368: \frac{{\cal P}_m(\kappa r_0, \kappa d)}{{\cal P}(\kappa r_0)} .
1369: \end{equation}
1370: Figure~\ref{fig:vmen_scaling} shows indeed that for two choices of $\kappa r_0$ this approximation is
1371: reasonable at large $d$. Numerically we find that the
1372: term $\propto \epsf \epsm$ contributes ca.~$70\%$ of the total
1373: meniscus--induced potential $\vmen$.
1374: %
1375: Asymptotically for large $d$ it exhibits the predicted
1376: $1/d^3$--behavior (Eq.~(\ref{eq:vmen_asymp}) with $n=6$). More
1377: precisely, the asymptotic behavior of $V_{\rm men}$ can be obtained as
1378: an expansion in terms of $1/d$ by utilizing the two--peak structure of
1379: $\Pi_{\rm m}$ (as for the general discussion in
1380: Subsec.~\ref{sec:asymptotic}): One assumes that for $d\to\infty$
1381: the main contribution
1382: to the integral in Eq.~(\ref{eq:defPm}) stems from the
1383: regions near the colloidal particles. The leading asymptotic behavior
1384: of ${\cal P}_m$ is given simply by the lowest--order term of a
1385: Taylor--expansion about $x_1 \approx \kappa r_0, x_2 \approx \kappa d$
1386: and $x_1 \approx \kappa d, x_2 \approx \kappa r_0$ in the integral in
1387: Eq.~(\ref{eq:defPm}) with the asymptotic decay
1388: for the ${\cal I}$--functions obtained in
1389: Appendix~\ref{app:Ifunctions}\footnote{Next-to-leading terms in the expansion
1390: require a more
1391: elaborate calculation than the simple Taylor--expansion, which
1392: yields undefined expressions for them.}:
1393: \begin{equation}
1394: \label{eq:pmasy}
1395: {\cal P}_m (\kappa r_0, \kappa d) \approx
1396: \frac{\epsilon_1}{\epsilon_2}\frac{p_m(\kappa r_0)}{(\kappa d)^3} ,
1397: \end{equation}
1398: where
1399: \begin{eqnarray}
1400: p_m (\kappa r_0) & := & 2 \int_{\kappa r_0}^\infty dx\,
1401: \left[
1402: \left(1-\frac{\epsilon_1}{\epsilon_2}\right)
1403: \frac{{\cal I}_b(x)}{x} + {\cal I}_a(x)
1404: \right] \\
1405: & & \nonumber \\
1406: & \approx & 2 \left[
1407: \kappa r_0 \, I_0\left(\frac{\kappa r_0}{2}\right) \,
1408: K_1\left(\frac{\kappa r_0}{2}\right)
1409: - 1 + {\rm e}^{-\kappa r_0}
1410: \right] , \nonumber
1411: \end{eqnarray}
1412: and the second line is the leading (zeroth order) contribution in an expansion in terms of the small
1413: parameter $\epsilon_1/\epsilon_2$, obtained from applying
1414: Eqs.~(\ref{eq:Iaeps}, \ref{eq:Ibeps}). According to this expression,
1415: $p_m (\kappa r_0 \to 0) = 4$ and $p(\kappa r_0)$ decreases
1416: monotonously with an asymptotic decay $p(\kappa r_0 \to\infty) \sim
1417: 2/(\kappa r_0)$.
1418: %
1419: This result for ${\cal P}_m (\kappa r_0, \kappa d)$ means that the
1420: asymptotic $1/d^3$-decay is determined by the normal component
1421: $E_{z}(d)$ of the field and
1422: by the potential $\Phi(d)$ stemming from the osmotic pressure. There
1423: is no contribution from ${\bf E}_\parallel$ to leading order in $1/d$
1424: due to the geometrical factor ${\bf x}_1 \cdot {\bf x}_2$ appearing in
1425: Eq.~(\ref{eq:defPm}).
1426: %
1427: Finally, from Fig.~\ref{fig:vmen_scaling} we infer that if $\kappa
1428: r_0$ is large enough, the asymptotic decay breaks down as a reliable
1429: approximation and a minimum appears in $\vmen(d)$ at a separation $d$
1430: which is a few times $\kappa^{-1}$.
1431:
1432: \begin{figure}[t]
1433: \begin{center}
1434: \vspace*{0.5cm}
1435: \epsfig{file=figures/vmen4.eps,width=.5\textwidth}
1436: \caption{The dimensionless capillary--induced potential energy
1437: $\bar{V}_{\rm men} := 10^2 \vmen/ [\gamma r_0^2 \epsf^2/ {\cal P}
1438: (\kappa r_0)]$ for two different values of $\kappa r_0$ and the
1439: choice $\epsilon_2/\epsilon_1=81$. Thick lines correspond to the
1440: capillary--induced potential given by Eq.~(\ref{eq:Vcorr2})
1441: (see Eq.~(\ref{eq:vmen_asymp})), whereas thin
1442: lines show the approximation $\propto \epsf \epsm$ given by
1443: Eq.~(\ref{eq:vmen_debye}).
1444: %
1445: If $\kappa r_0$ is large enough, a minimum appears.
1446: }
1447: \label{fig:vmen_scaling}
1448: \end{center}
1449: \end{figure}
1450:
1451: The energy due to the direct repulsion of the two colloids is given by
1452: Eq.~(\ref{eq:Efields}a):
1453: \begin{equation}
1454: V_{\rm rep}(d) = q \Phi(d) =
1455: 4\pi \gamma r_0^2 \epsf
1456: \frac{{\cal I}_a(\kappa d)}{(\kappa d) (\kappa r_0)
1457: {\cal P}(\kappa r_0)} ,
1458: \end{equation}
1459: and the total energy is $V_{\rm tot}(d) = V_{\rm rep}(d) + \vmen(d)$.
1460: With the approximation~(\ref{eq:vmen_debye}) one has
1461: \begin{equation}
1462: \label{eq:vtot_debye}
1463: V_{\rm tot}(d) \approx
1464: \frac{4\pi \gamma r_0^2 \epsf}{\kappa r_0\,{\cal P}(\kappa r_0)}
1465: \left[
1466: \frac{{\cal I}_a(\kappa d)}{\kappa d}
1467: - \epsf \, \kappa r_0 \, {\cal P}_m(\kappa r_0,\kappa d)
1468: \right] .
1469: \end{equation}
1470: Asymptotically for $\kappa d \gg 1$ this expression reduces to (see
1471: Eqs.~(\ref{eq:pmasy}, \ref{eq:iaasy}))
1472: \begin{equation}
1473: \label{eq:Vkappa_asymp}
1474: V_{\rm tot}(d) \sim
1475: 4\pi \gamma r_0^2 \epsf
1476: \frac{\epsilon_1/\epsilon_2}{(\kappa r_0)^4\,{\cal P}(\kappa r_0)}
1477: \left(\frac{r_0}{d}\right)^3
1478: \left[ 1
1479: + \frac{\epsilon_2}{\epsilon_1} (\kappa d)^2 \,
1480: {\rm e}^{-\kappa d}
1481: - \kappa r_0 \, p_m(\kappa r_0) \, \epsf
1482: \right] ,
1483: \end{equation}
1484: to be compared with the potential obtained in the previous subsection~(Eq.~(\ref{eq:Vtotal})). As in that case, the capillary--induced
1485: potential $\vmen$ is reduced by a factor $\epsf \ll 1$ with respect to
1486: $V_{\rm rep}$, and the total potential can be asymptotically
1487: attractive only if $\epsf$ is above a critical value
1488: \begin{equation}
1489: \label{eq:approx_critF}
1490: \varepsilon_{F,\rm{crit}} = \frac{1}{\kappa r_0 \, p_m(\kappa r_0)} ,
1491: \end{equation}
1492: which to leading order is independent of the small ratio
1493: $\epsilon_1/\epsilon_2$ and turns out to be bounded as
1494: $\varepsilon_{F,\rm{crit}} > 0.38$ (see Fig.~\ref{fig:critF}).
1495: %
1496: Moreover, we note that even if $\epsf \sim 1$ the term in brackets in
1497: Eq.~(\ref{eq:Vkappa_asymp}) changes sign at a separation $d$
1498: comparable with the crossover length $r_{\rm cross}$ (see
1499: Eq.~(\ref{eq:rcross})). This is a consequence of the crossover in
1500: $V_{\rm rep}(d)\propto \Phi(d)$ and suggests that although the
1501: potential is asymptotically attractive, it reaches a minimum at a
1502: distance $d\approx r_{\rm cross}$ and turns repulsive for closer
1503: separations.
1504: %
1505: This effect can only be enhanced
1506: by the deviation from the asymptotic $1/d^3$--decay in $V_{\rm
1507: men}(d)$ observed in Fig.~\ref{fig:vmen_scaling} for $\kappa r_0
1508: \sim 1$.
1509:
1510: These conclusions based on the approximation~(\ref{eq:vmen_debye}) are
1511: supported by the corresponding full numerical calculations.
1512: More precisely, from a fit to these numerical results
1513: in the
1514: range $0.1 \leq \kappa r_0 \leq 2.0$ for the ratio
1515: $\epsilon_2/\epsilon_1 = 81$ we find that the critical value is given
1516: approximately by
1517: \begin{equation}
1518: \label{eq:fit_critF}
1519: \varepsilon_{F, \rm crit} (\kappa r_0) \approx
1520: 0.34 \sqrt{1+\frac{0.30}{(\kappa r_0)^2}} \, .
1521: \end{equation}
1522: Figure~\ref{fig:vtot_debye} shows a plot of $V_{\rm tot}(d)$ for a
1523: typical value $\epsf =0.6$, exhibiting a shallow minimum
1524: if it is asymptotically attractive.
1525: %
1526: In summary, the capillary--induced attraction can dominate asymptotically
1527: only if $\kappa r_0$ is sufficiently large so that
1528: $\epsf>\varepsilon_{F,{\rm crit}}$, in which case one necessarily has
1529: $\epsf \gtrsim 1$.
1530:
1531: \begin{figure}[t]
1532: \begin{center}
1533: \vspace*{0.5cm}
1534: \epsfig{file=figures/critF.eps,width=.5\textwidth}
1535: \caption{The solid line gives the critical value $\varepsilon_{F,
1536: \rm crit}$ as a function of $\kappa r_0$ specified by the fit
1537: to the numerical results (Eq.~(\ref{eq:fit_critF})). The dashed
1538: line corresponds to the approximate expression provided by
1539: Eq.~(\ref{eq:approx_critF}).}
1540: \label{fig:critF}
1541: \end{center}
1542: \end{figure}
1543:
1544: \begin{figure}[t]
1545: \begin{center}
1546: \epsfig{file=figures/vtot_debye.eps,width=.5\textwidth}
1547: \caption{ The dimensionless total potential energy $\bar{V}_{\rm
1548: tot} := 10^3 V_{\rm tot} / [\gamma r_0^2 \epsf / {\cal P}
1549: (\kappa r_0)]$
1550: for different values of $\kappa r_0$ and the choices
1551: $\epsilon_2/\epsilon_1=81$ and $\epsf=0.6$. Note that the total
1552: energy is about 10 times smaller than $\vmen$
1553: (the normalization of $\bar{V}_{\rm tot}$ here differs from
1554: that of $\bar{V}_{\rm men}$ in Fig.~\ref{fig:vmen_scaling}
1555: by a factor $10/\epsf$)
1556: and exhibits a
1557: shallow minimum provided that $\kappa r_0$ is not too small.}
1558: \label{fig:vtot_debye}
1559: \end{center}
1560: \end{figure}
1561:
1562:
1563:
1564:
1565: \section{Summary and Discussion}
1566: \label{sec:end}
1567:
1568: We have carried out a thorough analysis of the effective interaction
1569: potential $\vmen(d)$ between two colloidal particles floating a
1570: distance $d$ apart from each other at a fluid interface due to the deformation of the
1571: interface caused by the particles. The main result is summarized in
1572: Eq.~(\ref{eq:vmen_asymp}) for the asymptotic behavior of $\vmen(d)$ as
1573: $d\to\infty$. One finds two qualitatively different cases depending on
1574: the (dimensionless) force $\epsf$ acting on the particle, and the
1575: one, $\epsp$,
1576: acting on the interface (see Eqs.~(\ref{eq:epsfp},
1577: \ref{eq:additiveF})): (a) if the system is not mechanically isolated
1578: ($\epsf-\epsp \neq 0$), the superposition approximation is valid and
1579: the asymptotic dependence $\vmen(d\to\infty)$ is universal in the
1580: sense of being independent of the detailed distribution of the forces
1581: which deform the interface. The physical reason for this is that each
1582: particle together with its surrounding interface can be considered as
1583: an effective particle.
1584: %
1585: (b) In the opposite case of mechanical isolation ($\epsf-\epsp = 0$),
1586: the corrections to the superposition ansatz are dominant and the
1587: asymptotic decay of $\vmen(d)$ depends on the asymptotic properties of
1588: the pressure field acting on the interface. We identified two sources
1589: of violation of the superposition approximation: the boundary
1590: conditions of the deformation at the particle--interface contact line
1591: are violated, and the pressure field acting on the interface does not
1592: satisfy a superposition principle.
1593:
1594: These results hold under the rather general assumptions that (i) the
1595: interfacial deformation is small ($\epsf, \epsp \ll 1$),
1596: (ii) the pressure field $\hat{\Pi} = \Pi_1 + \Pi_2 + 2 \Pi_m$ exerted by the
1597: particles on the interface satisfies the scaling relations in
1598: Eqs.~(\ref{eq:scalingPim}, \ref{eq:Pidecay}), and (iii) the external
1599: force acting on the particles is additive (Eq.~(\ref{eq:additiveF})).
1600: Assumptions (i) and (iii) are quite general and there is no evidence
1601: that experimentally they are not fulfilled. Assumption (ii) can be viewed as
1602: a consequence of the condition that the pressure $\hat{\Pi}$ is
1603: derivable from the stress tensor of a Lagrangian quadratic in an
1604: underlying field, as it is typically the case for physical systems:
1605: Besides stresses induced by electric fields as addressed here, there can be, e.g.,
1606: elastic stresses which arise if one of the fluids is in a nematic phase
1607: \cite{ODTD07}. (This latter case provides also an example in which the asymptotic
1608: decay~(\ref{eq:vmen_asymp}) is actually modified by geometrical
1609: constraints which complement the scaling
1610: relation~(\ref{eq:scalingPim}).)
1611:
1612: As an application for the case of mechanical isolation, we have
1613: considered the paradigmatic system of electrically charged colloids
1614: (see Figs.~\ref{fig:2coll_side} and~\ref{fig:2coll_top}), which is also of direct experimental
1615: importance. The pressure deforming the interface is due to the
1616: electrostatic field emanating from the particles and gives rise to an
1617: effective attractive interaction $\vmen(d\to\infty) \sim 1/d^3$.
1618: Since the direct electric repulsion between the colloids also decays
1619: $V_{\rm rep} \sim 1/d^3$, this asymptotic analysis is insufficient to
1620: determine whether the total potential $\vmen + V_{\rm rep}$ describes
1621: an asymptotically attractive force as seemingly observed
1622: experimentally. To this end, as a model for the experiments it is necessary to consider in detail the
1623: challenging electrostatic problem posed by two charged spheres
1624: partially immersed in an electrolyte.
1625: Within this approach we have studied the electrostatic problem in two
1626: simplifying limiting regimes: (A) water as one of the fluid phases is
1627: a perfect conductor, and (B) the colloidal particles are replaced by
1628: point charges (monopolar approximation).
1629: %
1630: Within both approximations we concluded that $\vmen \propto \epsf^2$,
1631: while $V_{\rm rep} \propto \epsf$, so that the direct repulsion
1632: dominates asymptotically; the capillary--induced attraction is
1633: predicted to dominate only if $\epsf$ is larger than a threshold value
1634: $\varepsilon_{F,{\rm crit}} \approx 1$ (see Eqs.~(\ref{eq:critical}, \ref{eq:fit_critF})), i.e., the deformation of the
1635: interface has to be large,
1636: which is outside the range of validity of the present analysis based on small
1637: deformations ($\epsf\ll 1$).
1638:
1639: Assumption (A) is a simple model of the experimentally relevant case
1640: of polystyrene or glass colloids floating at the interface between
1641: salty water and air or oil such that the Debye length $\kappa^{-1}$
1642: of water is much smaller than the radius of the particle--interface
1643: contact line $r_0$. Only the residual charges at the interface between
1644: the colloid and the air or oil phase determine $\epsf$. The charge
1645: density $\sigma$ at this interface is rather small compared with the nominal surface density of dissociable groups; we note that in
1646: the case of polystyrene colloids the precise microscopic origin of
1647: this charge density is still unknown. By dimensional analysis, $\epsf
1648: = (\sigma^2\,r_0/\epsilon_1 \gamma) \, G(\epsilon_{\rm
1649: C}/\epsilon_1,\theta)$, where $G$ is a dimensionless function
1650: depending on the contact angle $\theta$ and the ratio of dielectric
1651: constants of the colloidal particle and the insulating fluid,
1652: $\epsilon_{\rm C}$ and $\epsilon_1$, respectively. The
1653: electrostatic solution \cite{DaKr06a} yields $G(1,\theta) \approx
1654: 3/\sin^3\theta$ for hydrophobic colloids.
1655: %
1656: In \rcite{DKB06}\footnote{This work is {\it de facto} a correction
1657: \cite{ODD06} of previous work \cite{DKB04}.} the single--colloid
1658: meniscus deformation was measured around rather large glass spheres
1659: ($r_0 \approx 200\, \mu$m) at a water--oil interface ($\gamma \approx
1660: 0.05$ N/m, $\epsilon_1\approx 2$), which are slightly hydrophobic
1661: ($\theta \approx 120^o$). Using Eq.~(\ref{eq:single_u}) a value
1662: $\epsf \sim 0.4$ was inferred from the measurements, corresponding to
1663: a charge density of $ \sigma \sim 70\, \mu$C/m$^2$ (or $5\times
1664: 10^{-4}$ $e$/nm$^2$) according to the simple
1665: formula given above. This value is close to $\varepsilon_{F,{\rm
1666: crit}}$ given in Eq.~(\ref{eq:critical}) and thus one could expect
1667: a strongly reduced repulsion or even a net attraction between pairs of
1668: these glass spheres. A corresponding extension of this single--sphere experiment
1669: would be highly desirable.
1670:
1671: For truly nanoscopic colloids ($r_0 \alt 1\,\mu$m) with this same
1672: charge density on the air or oil side, the formula above predicts
1673: $\epsf =O(10^{-3})$ and therefore the electrostatic repulsion would
1674: always dominate the capillary attraction. This is consistent with the
1675: results in \rcite{ABCF02a} obtained for polystyrene spheres
1676: ($r_0\approx 1\,\mu$m) at the oil/water interface. For highly salty
1677: water, the charge density on the colloid/oil interface is estimated
1678: experimentally to be $\sigma \sim 20\, \mu$C/m$^2$ and force
1679: measurements between two spheres confirmed the repulsive
1680: dipole--dipole interaction with no sign of capillary attraction.
1681:
1682: Assumption (B) amounts to modelling a system in which the Debye length of water is comparable
1683: or larger than the radius $r_0$. We have studied in detail the case
1684: that the dielectric constant of one fluid phase (e.g., water) is much
1685: larger than the dielectric constant of the other fluid phase (e.g.,
1686: air).
1687: %
1688: The capillary--induced potential $\vmen$ is asymptotically attractive
1689: (see Fig.~\ref{fig:vmen_scaling}) but the total potential $\vmen +
1690: V_{\rm rep}$ can be asymptotically attractive
1691: for values $\epsf \sim 1$ only if
1692: $\kappa r_0 \gtrsim 1$ (see Fig.~\ref{fig:critF}). However, unlike
1693: the model corresponding to assumption (A), even in such a case
1694: $\vmen + V_{\rm rep}$ becomes repulsive at small separations
1695: (Fig.~\ref{fig:vtot_debye}). This effect
1696: can be traced back to a crossover in the interfacial stress
1697: $\Pi (r)$ from the asymptotic algebraic decay (as in model (A)) to an
1698: exponential decay at closer distances $r \lesssim 7/\kappa$ from the
1699: particle (see Fig.~\ref{fig:cross}).
1700:
1701: In contrast to the experiments carried out with nano-- and
1702: microcolloids at interfaces with salty water,
1703: some experiments \cite{QMMH01,GEFM01,NBHD02,TAKK03,GoRu05}
1704: have been performed with microcolloids at interfaces of ultrapure water
1705: such that $\kappa^{-1} \sim r_0 \sim 1$ $\mu$m.
1706: These experiments could have explored phenomena beyond the small deformation regime.
1707: Equation~(\ref{eq:qeff}) provides a relationship between $\epsf$ and
1708: the relevant parameters of the experimental system; regretably, the value of the total charge $q$ is
1709: usually uncertain. In terms of the surface charge density $\sigma$ one
1710: has $q=2\pi\sigma (r_0 \sin\theta)^2 (1+\cos\theta)$, because for ultrapure water the
1711: electrostatic field is dominated by the unscreened charge of the
1712: particle on the water side. For typical values $\epsilon_2=81$,
1713: $\gamma = 0.05\,$N/m, $\kappa = 1\,\mu$m$^{-1}$,
1714: and $\theta=\pi/2$ (so that $r_0=$ radius of the particle),
1715: Eq.~(\ref{eq:qeff}) gives $\epsf \approx 371 \, \sigma^2 \, r_0^3 \,
1716: {\cal P}(r_0)$ with $r_0$ in $\mu$m and $\sigma$ in units of
1717: $e$/nm$^2$. The values for $\sigma$ quoted in the
1718: literature range from $0.07\,e/$nm$^2$ (\rcite{GhEa97}) to
1719: $0.53\,e/$nm$^2$ (\rcite{SDJ00}). Accordingly, Fig.~(\ref{fig:modelB}) shows
1720: that it seems possible to have capillary attraction (i.e., $\epsf
1721: > \varepsilon_{F,{\rm crit}} \approx 1$) for typical values of the particle radius in the
1722: micrometer range. However, we emphasize that, unlike the conclusion
1723: concerning the asymptotic decay of $\vmen$, the expressions relating the value
1724: of $\epsf$ and $\varepsilon_{F,{\rm crit}}$ with the parameters of the
1725: system involve the behavior of the electric field near the particle
1726: (see Eqs.~(\ref{eq:calP}) and (\ref{eq:pmasy})). Therefore, they are
1727: expected to be affected by corrections to the monopolar approximation
1728: and the Debye--H{\"u}ckel approximation \cite{FDO07}. Thus, further theoretical
1729: work is required to understand these experiments properly and to
1730: arrive at reliable predictions.
1731:
1732: \begin{figure}[t]
1733: \begin{center}
1734: \epsfig{file=figures/modelB.eps,width=.4\textwidth}
1735: \caption{Parameter space spanned by the surface charge density $\sigma$
1736: (in units of $e$/nm$^2$) and the contact line radius $r_0$ (in
1737: $\mu$m) (see text for the fixed values of the other parameters). The solid line corresponds to the loci
1738: $\epsf=\varepsilon_{F,{\rm crit}}$, so that capillary attraction
1739: is predicted to occur in systems the parameters of which fall into the region
1740: above this curve. As a reference curve, the dashed line corresponds to the loci
1741: $\epsf=1$.}
1742: \label{fig:modelB}
1743: \end{center}
1744: \end{figure}
1745:
1746:
1747: \acknowledgments
1748:
1749: A.D. acknowledges financial support from the Junta de Andaluc{\'\i}a (Spain).
1750: %
1751: M.O. acknowledges financial support from the German Science Foundation
1752: (DFG) through the Collaborative Research Centre "Colloids in External
1753: Fields" (SFB--TR6).
1754:
1755: \appendix
1756:
1757: \section{Interface deformation fields}
1758: \label{app:ufunctions}
1759:
1760: In this appendix we derive the single--colloid deformation $u(r)$ and
1761: the correction $u_m(\br)$ to the superposition approximation for two
1762: colloids.
1763:
1764: The deformation field $u(r)$ in the single--colloid configuration is
1765: readily obtained as the rotationally symmetric solution of
1766: Eq.~(\ref{eq:single}):
1767: \begin{align}
1768: u(r) &= r_0 (\epsp - \epsf) \ln{\frac{L}{r}} -
1769: \frac{1}{\gamma} \int_{r}^{L} ds \; s \,
1770: \Pi(s) \, \ln{\frac{s}{r}} \nonumber \\
1771: \label{eq:single_u}
1772: &\sim \left\{
1773: \begin{aligned}
1774: r^{2-n}
1775: & \quad \textrm{if } \epsf-\epsp = 0 , \\
1776: \ln r
1777: & \quad \textrm{if } \epsf-\epsp \neq 0 ,
1778: \end{aligned}
1779: \right.
1780: \qquad (r_0 \ll r \ll L \textrm{ and } n>2).
1781: \end{align}
1782: The limit $L\to\infty$ is well defined if $\epsf-\epsp=0$; otherwise, the
1783: presence of the boundary condition~(\ref{eq:single_pin}) is required
1784: to regularize the possible logarithmic divergence \cite{ODD05a}.
1785: The physical interpretation of this regularization is a force acting on the boundary $C_L$ of the
1786: interface which compensates the net force $2\pi\gamma r_0
1787: (\epsf - \epsp)$ localized around the particle. Since these two forces
1788: act at well separated locations, there is an intermediate
1789: range of lengths where there is approximately no force acting on the
1790: interface, so that the corresponding deformation varies logarithmically.
1791: (The electrostatic analogy developed in \rcite{DOD06b} provides a
1792: transparent visualization of this explanation.)
1793:
1794: The correction $u_m(\br)$ to the superposition approximation cannot be
1795: computed analytically in an easy manner. But we shall derive several
1796: asymptotic properties of the solution as the interparticle
1797: separation $d$ becomes large. In order to follow the arguments,
1798: the reader will find the electrostatic analogy useful, which relies on
1799: the formal analogy of Eq.~(\ref{eq:cross}) with the equations for the
1800: two--dimensional electrostatic potential: $u_m$ plays the role of the
1801: potential and $2\Pi_m$ the role of the charge density. (This analogy
1802: is worked out in detail in \rcite{DOD06b}.)
1803:
1804: \begin{figure}
1805: \begin{center}
1806: \epsfig{file=figures/nearcolloid.eps,width=.3\textwidth}
1807: \caption{The position of a point $P$ near a colloid is parametrized
1808: by the polar coordinates $(\rho,\varphi)$. The unit vector
1809: normal to the (circular) contact line is $\bn = {\bf e}_\rho$.}
1810: \label{fig:um_near}
1811: \end{center}
1812: \end{figure}
1813:
1814: First we approximately compute the function $u_m(\br)$ near the
1815: colloids, i.e., at distances $r_0 \leq \rho := |\br - \br_\alpha| \ll d$
1816: from the particle (see Fig.~\ref{fig:um_near}). Concerning the
1817: boundary condition~(\ref{eq:cross_bc}) at $\rho=r_0$, we note that
1818: \begin{align}
1819: - \bn_\alpha \cdot \nabla u_\beta
1820: + \frac{u_\beta - \langle u_\beta \rangle}{r_0} & \approx
1821: \varepsilon_{12} - \frac{r_0}{2} \left[
1822: \bn \bn : \nabla\nabla u(d) - \frac{1}{2} \nabla^2 u(d)
1823: \right] \nonumber \\
1824: & \approx
1825: \varepsilon_{12} -
1826: \frac{r_0}{4} \left[ u''(d) - \frac{u'(d)}{d} \right] \cos 2\varphi ,
1827: \end{align}
1828: which follows from expanding the single--particle solution~(\ref{eq:single_u}) around
1829: $r=d$, where one finds (see Eq.~(\ref{eq:single_u}))
1830: \begin{equation}
1831: u''(d) - \frac{u'(d)}{d} \sim
1832: \left\{
1833: \begin{array}[c]{ccc}
1834: d^{-n}, & \quad \textrm{if } \epsf - \epsp = 0, \\
1835: d^{-2}, & \quad \textrm{if } \epsf - \epsp \neq 0 .
1836: \end{array}\right.
1837: \end{equation}
1838: %
1839: The field $\Pi_m(\br)$, according to the model discussed at the
1840: beginning of Subsec.~(\ref{sec:asymptotic}), is peaked at
1841: the colloids and decays far from them, so that it has approximately
1842: rotational symmetry in the range $r_0 < \rho \ll d$. More precisely,
1843: since $\Pi_m$ is proportional to $|\Pi(|\rho {\bf e}_\rho + d{\bf e}_x|)
1844: \Pi(\rho)|^{1/2}$, one finds, after expanding in terms of $\rho/d\ll 1$
1845: and in view of Eq.~(\ref{eq:Pidecay}),
1846: \begin{equation}
1847: \Pi_m(\br) \sim |\Pi(d) \Pi(\rho)|^{1/2} + {\cal O}(d^{-n/2-1}) .
1848: \end{equation}
1849: %
1850: Accordingly, in leading order in $1/d$ Eq.~(\ref{eq:cross}) turns into
1851: \begin{subequations}
1852: \begin{align}
1853: \frac{1}{\rho}\frac{\partial}{\partial \rho} \left[
1854: \rho \frac{\partial u_m}{\partial \rho} \right]
1855: +\frac{1}{\rho^2} \frac{\partial^2 u_m}{\partial \varphi^2}
1856: &\approx
1857: - \frac{2}{\gamma} {\Pi}_m (\rho) \sim d^{-n/2},
1858: \qquad r_0 < \rho \ll d , \\
1859: \label{eq:um_at_r0}
1860: \left[ \bn \cdot \nabla u_m
1861: - \frac{u_m - \langle u_m \rangle}{r_0} \right]_{\rho=r_0} & \approx
1862: \left\{
1863: \begin{array}[c]{cc}
1864: \epsm \sim d^{-n/2} , & \quad \textrm{if } \epsf - \epsp = 0 , \\
1865: - (r_0/4) [ u''(d) - u'(d)/d] \cos 2\varphi \sim d^{-2}, &
1866: \quad \textrm{if } \epsf - \epsp \neq 0 .
1867: \end{array}\right.
1868: \end{align}
1869: \end{subequations}
1870: Thus, if $\epsf-\epsp=0$ the near--particle solution is approximately
1871: rotationally symmetric and dominated by the pressure field $\Pi_m$:
1872: \begin{subequations}
1873: \label{eq:near_um}
1874: \begin{eqnarray}
1875: \label{eq:near_isol}
1876: u_m (\rho) & \approx &
1877: A_0 + r_0 \epsm \ln\frac{\rho}{r_0}
1878: + \frac{2}{\gamma} \int_{r_0}^{\rho} \!\!\!\! ds \; s \,
1879: \Pi_m (s) \ln \frac{s}{\rho}, \qquad \textrm{if } \epsf-\epsp=0 \nonumber \\
1880: & \approx &
1881: A_0 + \frac{2}{\gamma} \int_{r_0}^{\infty} \!\!\!\! ds \; s \,
1882: \Pi_m (s) \ln \frac{s}{r_0}
1883: + \frac{2}{\gamma} \int_{\rho}^\infty \!\!\!\! ds \; s \,
1884: \Pi_m (s) \ln \frac{\rho}{s},
1885: \end{eqnarray}
1886: where the second expression follows by inserting the
1887: estimate~(\ref{eq:approx_epsm}) for $\epsm$. Only the last term
1888: depends on $\rho$, while the second one is an additive constant
1889: proportional to $\sqrt{\Pi(d)} \sim d^{-n/2}$.
1890: %
1891: On the other hand, if $\epsf-\epsp \neq 0$ the near--particle solution
1892: is dominated by the single--particle solution for the boundary
1893: condition:
1894: \begin{equation}
1895: \label{eq:near_nonisol}
1896: u_m (\rho) \approx A_0 + \left\{
1897: \left[
1898: - \frac{r_0^2}{12} \left(\frac{u'(d)}{d} - u''(d) \right)
1899: + \frac{1}{3} A_2
1900: \right] \left(\frac{r_0}{\rho}\right)^2
1901: + A_2 \left(\frac{\rho}{r_0}\right)^2
1902: \right\} \cos 2\varphi , \quad \textrm{if } \epsf-\epsp \neq 0 .
1903: \end{equation}
1904: \end{subequations}
1905: The integration constants $A_0$ and $A_2$ are determined by the
1906: solution far from the particles and for our purposes here, it suffices
1907: to provide an estimate of how they depend on the separation $d$.
1908: In view of the electrostatic analogy the
1909: solution $u_m$ far from the particles, $r\gg d$, can be expressed
1910: in terms of a multipolar expansion. The
1911: ``capillary monopole''
1912: \begin{equation}
1913: \label{eq:cross_mono}
1914: Q_m := \int_{\hsmen} \!\!\!\!\! dA \; 2 \Pi_m +
1915: \gamma \oint_{\partial S_1 \bigcup \partial S_2} \!\!\! d\ell \;
1916: \bn\cdot (\mbox{}-\nabla u_m)
1917: \end{equation}
1918: is found to vanish exactly by virtue of the boundary condition~(\ref{eq:cross_bc}). The ``capillary dipole''
1919: vanishes due to the reflection symmetries
1920: of the configuration shown in Fig.~\ref{fig:2coll_top}
1921: upon $X\to\mbox{}-X$ and $Y\to\mbox{}-Y$.
1922: But in general the ``capillary quadrupole'' ${\mathsf D}_m$
1923: will be nonzero. Therefore the distant field is ($C$
1924: is a proportionality constant)
1925: \begin{equation}
1926: u_m (r \gtrsim d) \sim \frac{{\mathsf d}_m}{r^2} +
1927: C \ln\frac{r}{r_0} \int_{r}^{\infty} \!\!\!\! ds \; s \,
1928: \Pi_m (s) ,
1929: \end{equation}
1930: where we have neglected the angular dependence of the quadrupolar
1931: field (we are interested only in the decay with distance) and
1932: ${\mathsf d}_m$ denotes the typical value of the elements of the
1933: quadrupole ${\mathsf D}_m$.
1934: %
1935: The second term is approximately the ``potential'' created by the
1936: ``capillary charge'' beyond $r$, which is not accounted for by the multipolar moments,
1937: and arises as a correction to the
1938: multipolar expansion because the ``capillary charge'' $\Pi_m$ does not
1939: have a compact support. It scales as $r^{2-n} \ln r$. On the other
1940: hand, because the ``capillary monopole'' of a single particle is of the
1941: order of $\gamma r_0 \epsm$, the quadrupole of the ``capillary
1942: charge'' distributed over a region of size $\sim d$ will be ${\mathsf
1943: d}_m \sim d^2 \gamma r_0 \epsm \sim d^{2-n/2}$. Thus, the second
1944: term is negligible compared to the quadrupolar term if $r \sim d$,
1945: provided $n>4$, so that one finally finds $u_m (r\sim d) \sim
1946: d^{-n/2}$. For reasons of consistency we expect that the near--particle solutions
1947: in Eq.~(\ref{eq:near_um}) should scale like this if extrapolated to
1948: $\rho \sim d$: if Eq.~(\ref{eq:near_isol}) is evaluated at $\rho\sim
1949: d$ one obtains $A_0 ={\cal O}(d^{-n/2})$, while from
1950: Eq.~(\ref{eq:near_nonisol}) it follows that $A_2 = {\cal O}(d^{-4})$.
1951:
1952: In sum, the amplitude of the near--particle solution scales with
1953: the interparticle separation as follows:
1954: \begin{equation}
1955: \label{eq:cross_u}
1956: u_m (\rho \sim r_0) \sim \left\{
1957: \begin{aligned}
1958: d^{-n/2} ,
1959: & \quad \textrm{if } \epsf-\epsp = 0 , \\
1960: d^{-2} ,
1961: & \quad \textrm{if } \epsf-\epsp \neq 0 ,
1962: \end{aligned}
1963: \right. \qquad (n>4).
1964: \end{equation}
1965:
1966:
1967: \section{Flotation force and disjoining pressure}
1968: \label{app:flot}
1969:
1970: In this appendix we discuss briefly how our results are modified if
1971: the gravitational force is relevant. We also show how the same formal
1972: results hold if the lower fluid phase is a thin film on which, instead
1973: of gravity, dispersion forces due to a confining substrate are acting,
1974: exerting the so-called disjoining pressure.
1975:
1976: \subsection{Flotation force}
1977:
1978: The effect of the acceleration of gravity $g$ gives rise to an
1979: additional contribution to the free energy~(\ref{eq:freeF}): the
1980: gravitational potential energy of the fluids with respect to the
1981: reference configuration is
1982: \begin{equation}
1983: \label{eq:Fgrav}
1984: {\cal F}_{\rm grav} = \frac{1}{2} \gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
1985: \left(\frac{\hat{u}}{\lambda}\right)^2 ,
1986: \end{equation}
1987: where we have introduced the capillary length $\lambda :=
1988: \sqrt{\gamma/[(\varrho_--\varrho_+)g]}$ in terms of the mass densities
1989: $\varrho_+$ and $\varrho_-$ of the upper and lower fluid phases,
1990: respectively. This length has typical values in the millimeter range.
1991: %
1992: A pressure field (force per unit area) can be associated with this free
1993: energy:
1994: \begin{equation}
1995: \label{eq:Pigrav}
1996: \hat{\Pi}_{\rm grav}(\br) := \mbox{} -
1997: \frac{\delta{\cal F}_{\rm grav}}{\delta\hat{u}(\br)} =
1998: \mbox{} - \frac{\gamma}{\lambda^2} \hat{u}(\br) .
1999: \end{equation}
2000: This expression can be inserted directly into Eq.~(\ref{eq:equil_YL}),
2001: and one finds that at large distances from the particles $\hat{u}(\br)
2002: \sim \exp(-r/\lambda)$ and the field $\hat{\Pi}_{\rm grav}(\br)$ does
2003: indeed decay sufficiently fast.
2004: %
2005: However, one cannot apply the results we have derived previously
2006: without certain changes because $\hat{\Pi}_{\rm grav}$ depends explicitly
2007: on $\hat{u}$. Nevertheless, in the limit of a large capillary length
2008: one can neglect ${\cal F}_{\rm grav}$ (and thus $\hat{\Pi}_{\rm grav}$)
2009: altogether and retain only the gravitational force acting directly on
2010: the colloidal particle, i.e., $\epsp=0$ and $\hat{F} = F$ is the
2011: weight of the colloidal particle (corrected for buoyancy effects).
2012: In this case $\vmen(d)$ is given by the superposition approximation (see
2013: Eq.~(\ref{eq:vsup_asymp})) and reproduces the flotation force in the
2014: regime $d\ll\lambda$ \cite{CHW81}.
2015:
2016:
2017: \subsection{Disjoining pressure}
2018:
2019: \begin{figure}
2020: \begin{center}
2021: \epsfig{file=figures/film.eps,width=.5\textwidth}
2022: \caption{Schematic drawing of a fluid film on top of a substrate.
2023: $\ell$ is the thickness of the film in the reference, flat
2024: configuration, and $\hat{u}$ is the deformation of the interface.}
2025: \label{fig:film}
2026: \end{center}
2027: \end{figure}
2028:
2029: If the size of the particle lies below the micrometer, gravity is
2030: quantitatively negligible \cite{ODD05a}. The same formalism, however,
2031: is applicable in the experimentally relevant case that the lower fluid
2032: phase is a film of thickness $h(\br) = \ell + \hat{u}(\br)$ on top of
2033: a (solid or liquid) substrate (see Fig.~\ref{fig:film}).
2034: %
2035: If $\ell$ is within the range of the underlying dispersion forces, an
2036: additional contribution ${\cal F}_{\rm disp}$ to the free
2037: energy~(\ref{eq:freeF}) arises \cite{Diet88} (neglecting a constant,
2038: $\hat{u}$--independent term):
2039: \begin{equation}
2040: {\cal F}_{\rm disp} = \int_{\hsmen} \!\!\!\!\!\! dA \;
2041: \left[ \frac{H}{(\ell + \hat{u})^2} +
2042: (\ell + \hat{u}) \Delta n\, \Delta \mu \right] ,
2043: \end{equation}
2044: where $H$ is known as the Hamaker constant,
2045: $\Delta n$ is the number density difference between the bulk phases
2046: the film and the upper phase belong to,
2047: and $\Delta \mu$ is the
2048: undersaturation of the upper phase in terms of the chemical potential. The
2049: condition that the flat film of thickness $\ell$ is an equilibrium solution
2050: imposes the relation $ \Delta n\,\Delta\mu = 2H/\ell^3$.
2051: %
2052: The corresponding pressure field associated with ${\cal F}_{\rm disp}$
2053: is called ``disjoining pressure'':
2054: \begin{equation}
2055: \hat{\Pi}_{\rm disp} (\br) = \mbox{} -
2056: \frac{\delta{\cal F}_{\rm disp}}{\delta\hat{u}(\br)} =
2057: \frac{2 H}{(\ell + \hat{u})^3} - \frac{2 H}{\ell^3} .
2058: \end{equation}
2059:
2060: In the regime of small deformations one has $|\hat{u}| \ll \ell$ and ${\cal
2061: F}_{\rm disp}$ can be expanded around $\hat{u} = 0$ yielding
2062: \begin{equation}
2063: {\cal F}_{\rm disp} \approx {\cal F}_{\rm disp}[\hat{u}=0] +
2064: \frac{1}{2} \gamma \int_{\hsmen} \!\!\!\!\!\! dA \;
2065: \left(\frac{\hat{u}}{\lambda_{\rm disp}}\right)^2 ,
2066: \end{equation}
2067: which has the same form as Eq.~(\ref{eq:Fgrav}) with an effective
2068: ``capillary length'' $\lambda_{\rm disp} = \sqrt{\gamma\ell^4/(6 H)}$
2069: (which actually coincides with the so-called lateral correlation
2070: length $\xi_\parallel$ \cite{Diet88}). Thus $\hat{\Pi}_{\rm
2071: disp}$ takes the form of Eq.~(\ref{eq:Pigrav}) with $\lambda$
2072: replaced by $\lambda_{\rm disp}$.
2073: %
2074: If the film is sufficiently thin, the length $\lambda_{\rm disp}$ can
2075: be so small that it becomes relevant and a phenomenology may arise
2076: which is similar to the macroscopic one induced
2077: by gravity. For typical values $H = 10^{-20}\,$J and $\gamma =
2078: 0.05\,$N/m, one has $\lambda_{\rm disp} = 1\,\mu$m for a film
2079: thickness $\ell = 0.033\,\mu$m.
2080:
2081:
2082: \section{Effective capillary potential energy in the case of an ideally conducting fluid}
2083: \label{app:ideal_calc}
2084:
2085: In this appendix we provide the mathematical steps leading to
2086: Eq.~(\ref{eq:ideal_int}) which is valid in the limit $d \gg r_0$ and from which the
2087: effective capillary potential energy is obtained.
2088:
2089: In order to compute $\epsm$, we insert the ansatz~(\ref{eq:Pim_ideal})
2090: into the definition~(\ref{eq:epsm_def}). In the limit $d\to\infty$, the
2091: main contribution to the integral stems from the regions around each
2092: particle so that
2093: \begin{equation}
2094: \epsm \approx \frac{2}{\gamma r_0} \sqrt{\Pi(d)} \int_{r_0}^{\infty}
2095: dr \; r \sqrt{\Pi(r)}.
2096: \end{equation}
2097: With the pressure field given by Eq.~(\ref{eq:fitPi}) this reduces to
2098: \begin{equation}
2099: \epsm \approx 2 b(\mu) \epsf \left(\frac{r_0}{d}\right)^3
2100: \int_1^\infty dx \, \sqrt{(x-1)^{\mu-1} x^{-(\mu+3)}} ,
2101: \end{equation}
2102: and Eq.~(\ref{eq:ideal_int}a) is obtained upon performing the
2103: integral (Eq.~3.191.2 in \rcite{GrRy94}).
2104:
2105: For the evaluation of the integral in Eq.~(\ref{eq:ideal_int}b) we
2106: use the solutions~(\ref{eq:single_u}) and~(\ref{eq:near_isol}) valid
2107: for $\epsf-\epsm=0$. With Eq.~(\ref{eq:fitPi}) this leads to
2108: \begin{eqnarray}
2109: \label{eq:approx_nablau}
2110: \nabla u = \frac{\er}{\gamma r} \int_{r}^{\infty} ds \; s \, \Pi(s)
2111: & = &
2112: \er b(\mu) \epsf \frac{r_0}{r}
2113: \int_{r/r_0}^\infty dx \; (x-1)^{\mu-1} x^{-(\mu+4)} \nonumber \\
2114: & = &
2115: \frac{1}{4} \er b(\mu) \epsf \left(\frac{r_0}{r}\right)^5
2116: \; _2F_1 \left(1-\mu,4;5;\frac{r_0}{r}\right)
2117: \end{eqnarray}
2118: and, with the ansatz~(\ref{eq:Pim_ideal}), to
2119: \begin{eqnarray}
2120: \label{eq:approx_nablaum}
2121: \nabla u_m \approx \frac{2\er}{\gamma r} \int_{r}^{\infty} ds \; s \,
2122: \sqrt{\Pi(d) \Pi(s)}
2123: & = &
2124: 2 \er b(\mu) \epsf \frac{r_0}{r} \left(\frac{r_0}{d}\right)^3
2125: \int_{r/r_0}^\infty dx \; (x-1)^{(\mu-1)/2} x^{-(\mu+3)/2}
2126: \nonumber \\
2127: & = &
2128: 2 \er b(\mu) \epsf \left(\frac{r_0}{d}\right)^3
2129: \left(\frac{r_0}{r}\right)^2
2130: \; _2F_1 \left(\frac{1-\mu}{2},1;2;\frac{r_0}{r}\right) .
2131: \end{eqnarray}
2132: (Concerning the last lines in Eqs.~(\ref{eq:approx_nablau}) and
2133: (\ref{eq:approx_nablaum}) see Eq.~3.194.2 in \rcite{GrRy94}.)
2134: This enables one to obtain
2135: \begin{eqnarray}
2136: \int_{S_{\rm men, 2}} \!\!\!\!\!\! dA \;
2137: (\nabla u_m) \cdot (\nabla u_2) & \approx &
2138: \pi b(\mu)^2 \epsf^2 \left(\frac{r_0}{d}\right)^3 \times \mbox{} \\
2139: & & \mbox{} \times \int_{r_0}^{\infty} \!\!\!\!\!\!dr \; r \,
2140: \left(\frac{r_0}{r}\right)^7
2141: \, _2F_1 \left(1-\mu,4;5;\frac{r_0}{r}\right)
2142: \, _2F_1 \left(\frac{1-\mu}{2},1;2;\frac{r_0}{r}\right) \nonumber
2143: \end{eqnarray}
2144: so that Eqs.~(\ref{eq:ideal_int}b, \ref{eq:Mfunction}) follow upon a
2145: change of variable in the integral.
2146:
2147: \section{Asymptotic behavior of the electrostatic potential and of the electric field}
2148: \label{app:Ifunctions}
2149:
2150: In this appendix we discuss the asymptotic behaviors of the functions
2151: ${\cal I}_n(k)$ defined in Eq.~(\ref{eq:Idef}). First we mention that
2152: at first sight the integrals may appear to be divergent due to the
2153: weak power law decay of the integrands, which are, however,
2154: oscillatory. Actually, the integrals are
2155: regularized by an exponential, so that for instance
2156: \begin{equation}
2157: {\cal I}_a(k) := \lim_{h\rightarrow 0} \int_0^\infty dx\;
2158: {\rm e}^{-h x} J_0(x)\;
2159: \frac{x}{
2160: \frac{\epsilon_1}{\epsilon_2} x + \sqrt{x^2+k^2}} ,
2161: \end{equation}
2162: reflecting the physical situation that the charge $q$ is positioned at
2163: a (dimensionless) height $h>0$ above the flat interface. In the
2164: following mathematical manipulations this regularization scheme is
2165: implied, which, unless required, we do not write explicitly to avoid a
2166: clumsy notation.
2167:
2168: Reference~\cite{Hurd85} provides a method to obtain the expansion of
2169: ${\cal I}_a(k)$ in powers of $\epsilon_1/\epsilon_2 \ll 1$.
2170: The idea is to split the integrals into a
2171: sum of two terms involving only odd or even powers of the ratio
2172: $\epsilon_1/\epsilon_2$, respectively. This leads to\footnote{We note
2173: (i) a misprint in Eq.~(10) in \rcite{Hurd85},
2174: because there the term $2^m$ should be in the numerator, and (ii)
2175: that the series provided in that Eq.~(10) can actually be carried
2176: out and is proportional to the error function.}
2177: \begin{equation}
2178: \label{eq:Iaeps}
2179: {\cal I}_a (k) = {\rm e}^{-k} +
2180: \frac{\epsilon_1}{\epsilon_2} \left[
2181: \frac{\pi}{2} k (I_0 (k) - {\bf L}_0 (k)) - 1
2182: \right] + {\cal O}(\epsilon_1/\epsilon_2)^2 ,
2183: \qquad k \ll (\epsilon_2/\epsilon_1)^2 ,
2184: \end{equation}
2185: in terms of the Bessel function $I_0$ and the Struve function ${\bf L}_0$
2186: \cite{GrRy94}. We note that the validity of this expression is
2187: restricted to sufficiently small values of $k$. However, for the
2188: typical values of the ratio $\epsilon_2/\epsilon_1$ occurring in the
2189: experiments so far, this does not impose any physically relevant
2190: constraint on $k$. The asymptotic behavior for large $k$ is
2191: \begin{equation}
2192: \label{eq:iaasy}
2193: {\cal I}_a (k) \approx {\rm e}^{-k} +
2194: \frac{\epsilon_1}{\epsilon_2} \frac{1}{k^2} ,
2195: \qquad 1 \ll k \ll (\epsilon_2/\epsilon_1)^2 .
2196: \end{equation}
2197: As a general result \cite{Hurd85}, the coefficients of the odd powers of
2198: $\epsilon_1/\epsilon_2$ decay algebraically for
2199: $k\gg 1$, while the coefficients of the even powers decay exponentially.
2200: %
2201: With Eq.~(\ref{eq:Idef}c) this leads to
2202: \begin{equation}
2203: {\cal I}_c (k) \approx k \, {\rm e}^{-k} +
2204: 3\frac{\epsilon_1}{\epsilon_2} \frac{1}{k^2} ,
2205: \qquad 1 \ll k \ll (\epsilon_2/\epsilon_1)^2 .
2206: \end{equation}
2207: %
2208: The procedure employed in \rcite{Hurd85} can be extended to analyze
2209: ${\cal I}_b(k)$. The radical is eliminated from the denominator of the
2210: integrand in Eq.~(\ref{eq:Idef}b) and the integral is split as follows:
2211: \begin{equation}
2212: \label{eq:I2split}
2213: {\cal I}_b (k) = \frac{\epsilon_2}{\epsilon_1}
2214: \int_0^\infty dx\; J_0(x)\;
2215: \frac{x\,(x^2+k^2)}{[1-(\epsilon_1/\epsilon_2)^2] x^2 + k^2}
2216: - \int_0^\infty dx\; J_0(x)\;
2217: \frac{x^2\,\sqrt{x^2+k^2}}{[1-(\epsilon_1/\epsilon_2)^2] x^2 + k^2} .
2218: \end{equation}
2219: The first term involves only odd powers of $\epsilon_1/\epsilon_2$,
2220: the second term only even powers.
2221: The first integral can be rewritten as
2222: \begin{equation}
2223: \frac{\epsilon_2}{\epsilon_1} \int_0^\infty dx\; J_0(x)\;
2224: \frac{x\,(x^2+k^2)}{[1-(\epsilon_1/\epsilon_2)^2] x^2 + k^2} =
2225: \frac{\epsilon_2/\epsilon_1}{1-(\epsilon_1/\epsilon_2)^2}
2226: \int_0^\infty dx\; J_0(x)\; x
2227: \left[ 1 + \frac{k^2-\frac{k^2}{[1-(\epsilon_1/\epsilon_2)^2]}}
2228: {x^2 + \frac{k^2}{[1-(\epsilon_1/\epsilon_2)^2]}} \right] .
2229: \end{equation}
2230: With the identities (see Eqs.~6.623.2 and~6.532.4 in \rcite{GrRy94})
2231: \begin{equation}
2232: \int_0^\infty dx\; J_0(x)\; x = 0
2233: \end{equation}
2234: and
2235: \begin{equation}
2236: \int_0^\infty dx\; J_0(x)\; \frac{x}{x^2 + a^2} = K_0(a)
2237: \end{equation}
2238: the first integral in Eq.~(\ref{eq:I2split}) can be written as
2239: \begin{eqnarray}
2240: \label{eq:Ibevenint}
2241: \frac{\epsilon_2}{\epsilon_1} \int_0^\infty dx\; J_0(x)\;
2242: \frac{x\,(x^2+k^2)}{[1-(\epsilon_1/\epsilon_2)^2] x^2 + k^2} & = &
2243: \mbox{}-\frac{\epsilon_1/\epsilon_2}{[1-(\epsilon_1/\epsilon_2)^2]^2}
2244: \; k^2 \; K_0 \left( k/\sqrt{1-(\epsilon_1/\epsilon_2)^2} \right)
2245: \nonumber \\
2246: & & \nonumber \\
2247: & = & \mbox{} - (\epsilon_1/\epsilon_2)
2248: \; k^2 \; K_0 (k) + {\cal O}(\epsilon_1/\epsilon_2)^3 .
2249: \end{eqnarray}
2250: The second integral in Eq.~(\ref{eq:I2split}) can be written to
2251: leading order in $\epsilon_1/\epsilon_2$ as
2252: \begin{equation}
2253: \label{eq:Ib2ndintegral}
2254: \int_0^\infty dx\; J_0(x)\;
2255: \frac{x^2\,\sqrt{x^2+k^2}}{x^2 + k^2} =
2256: \int_0^\infty dx\; J_0(x)\;
2257: \left[\sqrt{x^2+k^2} - \frac{k^2}{\sqrt{x^2 + k^2}}\right] .
2258: \end{equation}
2259: We introduce the exponential regularization as
2260: $\exp{(-h\sqrt{x^2+k^2})}$ and note the identity (Eq.~6.637.1 in
2261: \rcite{GrRy94})
2262: \begin{equation}
2263: \int_0^\infty dx\; J_0(x)\;
2264: \frac{{\rm e}^{-h\sqrt{x^2+k^2}}}{\sqrt{x^2 + k^2}} =
2265: I_0 \left(\frac{k}{2}\left[\sqrt{h^2+1}-h\right]\right) \,
2266: K_0 \left(\frac{k}{2}\left[\sqrt{h^2+1}+h\right]\right) .
2267: \end{equation}
2268: Therefore, the righthand side of Eq.~(\ref{eq:Ib2ndintegral}) can be rewritten as
2269: \begin{eqnarray}
2270: \label{eq:Iboddint}
2271: \int_0^\infty dx\; J_0(x)\;
2272: \frac{x^2\,\sqrt{x^2+k^2}}{x^2 + k^2} & = &
2273: \lim_{h\rightarrow 0} \left[
2274: \frac{\partial^2}{\partial h^2} - k^2
2275: \right]
2276: \int_0^\infty dx\; J_0(x)\;
2277: \frac{{\rm e}^{-h\sqrt{x^2+k^2}}}{\sqrt{x^2 + k^2}} \nonumber \\
2278: & & \nonumber \\
2279: & = & \frac{1}{2} k^2 \left[
2280: I_1 \left(\frac{k}{2}\right) \, K_1 \left(\frac{k}{2}\right)
2281: - I_0 \left(\frac{k}{2}\right) \, K_0 \left(\frac{k}{2}\right)
2282: \right] \\
2283: & & \nonumber \\
2284: & = & k \frac{d}{d k} \left[
2285: k \, I_0 \left(\frac{k}{2}\right) \, K_1 \left(\frac{k}{2}\right)
2286: \right] . \nonumber
2287: \end{eqnarray}
2288: Therefore from Eqs.~(\ref{eq:Ibevenint}) and~(\ref{eq:Iboddint}) one
2289: finally arrives at the following expansion:
2290: \begin{equation}
2291: \label{eq:Ibeps}
2292: {\cal I}_b(k) =
2293: \frac{1}{2} k^2 \left[
2294: I_0 \left(\frac{k}{2}\right) \, K_0 \left(\frac{k}{2}\right)
2295: - I_1 \left(\frac{k}{2}\right) \, K_1 \left(\frac{k}{2}\right)
2296: \right]
2297: - \frac{\epsilon_1}{\epsilon_2}
2298: \; k^2 \; K_0 (k) + {\cal O}(\epsilon_1/\epsilon_2)^3 .
2299: \end{equation}
2300: This expansion may also be restricted to sufficiently small values of
2301: $k$, although we have not found this upper bound as function of
2302: $\epsilon_1/\epsilon_2$.
2303: However, as argued above this possible constraint is expected to be
2304: physically irrelevant because typically $\epsilon_1/\epsilon_2$ is sufficiently
2305: small. The asymptotic behavior of this expression for $k\gg 1$ is
2306: \begin{equation}
2307: {\cal I}_b (k) \approx
2308: \frac{1}{k} - \sqrt{\frac{\pi}{2}} \;
2309: \frac{\epsilon_1}{\epsilon_2} \;
2310: k^{3/2}\, {\rm e}^{-k} ,
2311: \qquad 1\ll k, \, \epsilon_2/\epsilon_1.
2312: \end{equation}
2313:
2314: \newpage
2315: \begin{thebibliography}{35}
2316: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2317: \expandafter\ifx\csname bibnamefont\endcsname\relax
2318: \def\bibnamefont#1{#1}\fi
2319: \expandafter\ifx\csname bibfnamefont\endcsname\relax
2320: \def\bibfnamefont#1{#1}\fi
2321: \expandafter\ifx\csname citenamefont\endcsname\relax
2322: \def\citenamefont#1{#1}\fi
2323: \expandafter\ifx\csname url\endcsname\relax
2324: \def\url#1{\texttt{#1}}\fi
2325: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
2326: \providecommand{\bibinfo}[2]{#2}
2327: \providecommand{\eprint}[2][]{\url{#2}}
2328:
2329: \bibitem[{\citenamefont{Pieranski}(1980)}]{Pier80}
2330: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Pieranski}},
2331: \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{569}
2332: (\bibinfo{year}{1980}).
2333:
2334: \bibitem[{\citenamefont{Joannopoulos}(2001)}]{Joan01}
2335: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Joannopoulos}},
2336: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{414}},
2337: \bibinfo{pages}{257} (\bibinfo{year}{2001}).
2338:
2339: \bibitem[{\citenamefont{Dinsmore et~al.}(2002)\citenamefont{Dinsmore, Hsu,
2340: Nikolaides, M\'arquez, Bausch, and Weitz}}]{DHNM02}
2341: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dinsmore}},
2342: \bibinfo{author}{\bibfnamefont{M.~F.} \bibnamefont{Hsu}},
2343: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nikolaides}},
2344: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{M\'arquez}},
2345: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Bausch}}, \bibnamefont{and}
2346: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Weitz}},
2347: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{298}},
2348: \bibinfo{pages}{1006} (\bibinfo{year}{2002}).
2349:
2350: \bibitem[{\citenamefont{Hurd}(1985)}]{Hurd85}
2351: \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Hurd}},
2352: \bibinfo{journal}{J.\ Phys.\ A: Math.\ Gen.} \textbf{\bibinfo{volume}{18}},
2353: \bibinfo{pages}{L1055} (\bibinfo{year}{1985}).
2354:
2355: \bibitem[{\citenamefont{Aveyard et~al.}(2000)\citenamefont{Aveyard, Clint,
2356: Nees, and Paunov}}]{ACNP00}
2357: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Aveyard}},
2358: \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Clint}},
2359: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Nees}}, \bibnamefont{and}
2360: \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Paunov}},
2361: \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{16}},
2362: \bibinfo{pages}{1969} (\bibinfo{year}{2000}).
2363:
2364: \bibitem[{\citenamefont{Ruiz-Garc{\'\i}a
2365: et~al.}(1997)\citenamefont{Ruiz-Garc{\'\i}a, G\'amez-Corrales, and
2366: Ivlev}}]{RGI97}
2367: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ruiz-Garc{\'\i}a}},
2368: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{G\'amez-Corrales}},
2369: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Ivlev}},
2370: \bibinfo{journal}{Physica A} \textbf{\bibinfo{volume}{236}},
2371: \bibinfo{pages}{97} (\bibinfo{year}{1997}).
2372:
2373: \bibitem[{\citenamefont{Ghezzi and Earnshaw}(1997)}]{GhEa97}
2374: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ghezzi}} \bibnamefont{and}
2375: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Earnshaw}},
2376: \bibinfo{journal}{J.\ Phys.: Condensed\ Matt.} \textbf{\bibinfo{volume}{9}},
2377: \bibinfo{pages}{L517} (\bibinfo{year}{1997}).
2378:
2379: \bibitem[{\citenamefont{Stamou et~al.}(2000)\citenamefont{Stamou, Duschl, and
2380: Johannsmann}}]{SDJ00}
2381: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Stamou}},
2382: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Duschl}}, \bibnamefont{and}
2383: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Johannsmann}},
2384: \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{5263}
2385: (\bibinfo{year}{2000}).
2386:
2387: \bibitem[{\citenamefont{Quesada-P\'erez
2388: et~al.}(2001)\citenamefont{Quesada-P\'erez, Moncho-Jord\'a,
2389: Mart{\'\i}nez-L\'opez, and Hidalgo-Alvarez}}]{QMMH01}
2390: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Quesada-P\'erez}},
2391: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moncho-Jord\'a}},
2392: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Mart{\'\i}nez-L\'opez}},
2393: \bibnamefont{and}
2394: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hidalgo-Alvarez}},
2395: \bibinfo{journal}{J.\ Chem.\ Phys.} \textbf{\bibinfo{volume}{115}},
2396: \bibinfo{pages}{10897} (\bibinfo{year}{2001}).
2397:
2398: \bibitem[{\citenamefont{Ghezzi et~al.}(2001)\citenamefont{Ghezzi, Earnshaw,
2399: Finnis, and McCluney}}]{GEFM01}
2400: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ghezzi}},
2401: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Earnshaw}},
2402: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Finnis}}, \bibnamefont{and}
2403: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{McCluney}},
2404: \bibinfo{journal}{J.\ Coll.\ Interface Sci.} \textbf{\bibinfo{volume}{238}},
2405: \bibinfo{pages}{433} (\bibinfo{year}{2001}).
2406:
2407: \bibitem[{\citenamefont{Mej{\'\i}a-Rosales
2408: et~al.}(2002)\citenamefont{Mej{\'\i}a-Rosales, Gil-Villegas, Ivlev, and
2409: Ruiz-Garc{\'\i}a}}]{MGIR02}
2410: \bibinfo{author}{\bibfnamefont{S.~J.} \bibnamefont{Mej{\'\i}a-Rosales}},
2411: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Gil-Villegas}},
2412: \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Ivlev}}, \bibnamefont{and}
2413: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ruiz-Garc{\'\i}a}},
2414: \bibinfo{journal}{J. Phys.: Condensed Matt.} \textbf{\bibinfo{volume}{14}},
2415: \bibinfo{pages}{4795} (\bibinfo{year}{2002}).
2416:
2417: \bibitem[{\citenamefont{Nikolaides et~al.}(2002)\citenamefont{Nikolaides,
2418: Bausch, Hsu, Dinsmore, Brenner, Gay, and Weitz}}]{NBHD02}
2419: \bibinfo{author}{\bibfnamefont{M.~G.} \bibnamefont{Nikolaides}},
2420: \bibinfo{author}{\bibfnamefont{A.~R.} \bibnamefont{Bausch}},
2421: \bibinfo{author}{\bibfnamefont{M.~F.} \bibnamefont{Hsu}},
2422: \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Dinsmore}},
2423: \bibinfo{author}{\bibfnamefont{M.~P.} \bibnamefont{Brenner}},
2424: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gay}}, \bibnamefont{and}
2425: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Weitz}},
2426: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{420}},
2427: \bibinfo{pages}{299} (\bibinfo{year}{2002}).
2428:
2429: \bibitem[{\citenamefont{Tolnai et~al.}(2003)\citenamefont{Tolnai, Agod,
2430: Kabai-Faix, Kov{\'a}cs, Ramsden, and H{\'o}rv{\"o}lgyi}}]{TAKK03}
2431: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Tolnai}},
2432: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Agod}},
2433: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kabai-Faix}},
2434: \bibinfo{author}{\bibfnamefont{A.~L.} \bibnamefont{Kov{\'a}cs}},
2435: \bibinfo{author}{\bibfnamefont{J.~J.} \bibnamefont{Ramsden}},
2436: \bibnamefont{and}
2437: \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{H{\'o}rv{\"o}lgyi}},
2438: \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{107}},
2439: \bibinfo{pages}{11109} (\bibinfo{year}{2003}).
2440:
2441: \bibitem[{\citenamefont{G\'omez-Guzm\'an and Ruiz-Garc{\'\i}a}(2005)}]{GoRu05}
2442: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{G\'omez-Guzm\'an}}
2443: \bibnamefont{and}
2444: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ruiz-Garc{\'\i}a}},
2445: \bibinfo{journal}{J.\ Coll. and Interface Sci.}
2446: \textbf{\bibinfo{volume}{291}}, \bibinfo{pages}{1} (\bibinfo{year}{2005}).
2447:
2448: \bibitem[{\citenamefont{Chen et~al.}(2006)\citenamefont{Chen, Tan, Huang, Ng,
2449: Ford, and Tong}}]{CTHN06}
2450: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Chen}},
2451: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Tan}},
2452: \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Huang}},
2453: \bibinfo{author}{\bibfnamefont{T.-K.} \bibnamefont{Ng}},
2454: \bibinfo{author}{\bibfnamefont{W.~T.} \bibnamefont{Ford}}, \bibnamefont{and}
2455: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tong}},
2456: \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{74}},
2457: \bibinfo{pages}{021406} (\bibinfo{year}{2006}).
2458:
2459: \bibitem[{\citenamefont{Fern\'andez-Toledano
2460: et~al.}(2004)\citenamefont{Fern\'andez-Toledano, Moncho-Jord\'a,
2461: Mart{\'\i}nez-L\'opez, and Hidalgo-Alvarez}}]{FMMH04}
2462: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Fern\'andez-Toledano}},
2463: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moncho-Jord\'a}},
2464: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Mart{\'\i}nez-L\'opez}},
2465: \bibnamefont{and}
2466: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hidalgo-Alvarez}},
2467: \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{20}},
2468: \bibinfo{pages}{6977 } (\bibinfo{year}{2004}).
2469:
2470: \bibitem[{\citenamefont{Nicolson}(1949)}]{Nico49}
2471: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Nicolson}},
2472: \bibinfo{journal}{Proc.\ Cambridge Philos.\ Soc.}
2473: \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{288} (\bibinfo{year}{1949}).
2474:
2475: \bibitem[{\citenamefont{Chan et~al.}(1981)\citenamefont{Chan, {Henry Jr.}, and
2476: White}}]{CHW81}
2477: \bibinfo{author}{\bibfnamefont{D.~Y.~C.} \bibnamefont{Chan}},
2478: \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{{Henry Jr.}}},
2479: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{White}},
2480: \bibinfo{journal}{J.\ Coll.\ Interface Sci.} \textbf{\bibinfo{volume}{79}},
2481: \bibinfo{pages}{410} (\bibinfo{year}{1981}).
2482:
2483: \bibitem[{\citenamefont{Danov et~al.}(2004)\citenamefont{Danov, Kralchevsky,
2484: and Boneva}}]{DKB04}
2485: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}},
2486: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
2487: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~P.}
2488: \bibnamefont{Boneva}}, \bibinfo{journal}{Langmuir}
2489: \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{6139} (\bibinfo{year}{2004}).
2490:
2491: \bibitem[{\citenamefont{Megens and Aizenberg}(2003)}]{MeAi03}
2492: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Megens}} \bibnamefont{and}
2493: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Aizenberg}},
2494: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{424}},
2495: \bibinfo{pages}{1014} (\bibinfo{year}{2003}).
2496:
2497: \bibitem[{\citenamefont{Foret and W\"urger}(2004)}]{FoWu04}
2498: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Foret}} \bibnamefont{and}
2499: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W\"urger}},
2500: \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{92}},
2501: \bibinfo{pages}{058302} (\bibinfo{year}{2004}).
2502:
2503: \bibitem[{\citenamefont{Oettel et~al.}(2005{\natexlab{a}})\citenamefont{Oettel,
2504: Dom{\'\i}nguez, and Dietrich}}]{ODD05a}
2505: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
2506: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2507: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2508: \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{71}},
2509: \bibinfo{pages}{051401} (\bibinfo{year}{2005}{\natexlab{a}}).
2510:
2511: \bibitem[{\citenamefont{Dom{\'\i}nguez
2512: et~al.}(2005)\citenamefont{Dom{\'\i}nguez, Oettel, and Dietrich}}]{DOD05}
2513: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2514: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
2515: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2516: \bibinfo{journal}{J.\ Phys.: Condensed Matt.} \textbf{\bibinfo{volume}{17}},
2517: \bibinfo{pages}{S3387} (\bibinfo{year}{2005}).
2518:
2519: \bibitem[{\citenamefont{Oettel et~al.}(2006)\citenamefont{Oettel,
2520: Dom{\'\i}nguez, and Dietrich}}]{ODD06}
2521: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
2522: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2523: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2524: \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{22}},
2525: \bibinfo{pages}{846} (\bibinfo{year}{2006}).
2526:
2527: \bibitem[{\citenamefont{Oettel et~al.}(2005{\natexlab{b}})\citenamefont{Oettel,
2528: Dom{\'\i}nguez, and Dietrich}}]{ODD05b}
2529: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
2530: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2531: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2532: \bibinfo{journal}{J.\ Phys.: Condensed\ Matt.} \textbf{\bibinfo{volume}{17}},
2533: \bibinfo{pages}{L337} (\bibinfo{year}{2005}{\natexlab{b}}).
2534:
2535: \bibitem[{\citenamefont{W{\"u}rger and Foret}(2005)}]{WuFo05}
2536: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}} \bibnamefont{and}
2537: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Foret}}, \bibinfo{journal}{J.
2538: Phys. Chem. B} \textbf{\bibinfo{volume}{109}}, \bibinfo{pages}{16435}
2539: (\bibinfo{year}{2005}).
2540:
2541: \bibitem[{\citenamefont{Dom{\'\i}nguez
2542: et~al.}(2006)\citenamefont{Dom{\'\i}nguez, Oettel, and Dietrich}}]{DOD06b}
2543: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2544: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
2545: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2546: \bibinfo{journal}{\texttt{cond-mat/0611329}} (\bibinfo{year}{2006}).
2547:
2548: \bibitem[{\citenamefont{W{\"u}rger}(2006)}]{Wuer06b}
2549: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}},
2550: \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{978}
2551: (\bibinfo{year}{2006}).
2552:
2553: \bibitem[{\citenamefont{Dom{\'\i}nguez
2554: et~al.}(2007)\citenamefont{Dom{\'\i}nguez, Oettel, and Dietrich}}]{DOD07a}
2555: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2556: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
2557: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2558: \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{77}}, \bibinfo{pages}{68002}
2559: (\bibinfo{year}{2007}).
2560:
2561: \bibitem[{\citenamefont{Aveyard et~al.}(2002)\citenamefont{Aveyard, Binks,
2562: Clint, Fletcher, Horozov, Neumann, Paunov, Annesley, Botchway, Nees
2563: et~al.}}]{ABCF02a}
2564: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Aveyard}},
2565: \bibinfo{author}{\bibfnamefont{B.~P.} \bibnamefont{Binks}},
2566: \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Clint}},
2567: \bibinfo{author}{\bibfnamefont{P.~D.~I.} \bibnamefont{Fletcher}},
2568: \bibinfo{author}{\bibfnamefont{T.~S.} \bibnamefont{Horozov}},
2569: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Neumann}},
2570: \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Paunov}},
2571: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Annesley}},
2572: \bibinfo{author}{\bibfnamefont{S.~W.} \bibnamefont{Botchway}},
2573: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Nees}},
2574: \bibinfo{author}{\bibfnamefont{A.~W.}~\bibnamefont{Parker}},
2575: \bibinfo{author}{\bibfnamefont{A.~D.}~\bibnamefont{Ward}},
2576: \bibnamefont{and}
2577: \bibinfo{author}{\bibfnamefont{A.~N.}~\bibnamefont{Burgess}},
2578: %\bibnamefont{et~al.},
2579: \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{88}},
2580: \bibinfo{pages}{246102} (\bibinfo{year}{2002}).
2581:
2582: \bibitem[{\citenamefont{Danov and Kralchevsky}(2006)}]{DaKr06a}
2583: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}} \bibnamefont{and}
2584: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
2585: \bibinfo{journal}{J. Coll. Interface Sci.} \textbf{\bibinfo{volume}{298}},
2586: \bibinfo{pages}{213} (\bibinfo{year}{2006}).
2587:
2588: \bibitem[{\citenamefont{Gradshteyn and Ryzhik}(1994)}]{GrRy94}
2589: \bibinfo{author}{\bibfnamefont{I.~S.} \bibnamefont{Gradshteyn}}
2590: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{I.~M.}
2591: \bibnamefont{Ryzhik}}, \emph{\bibinfo{title}{Table of Integrals, Series, and
2592: Products}} (\bibinfo{publisher}{Academic}, \bibinfo{address}{London},
2593: \bibinfo{year}{1994}), \bibinfo{edition}{5th} ed.
2594:
2595: \bibitem[{\citenamefont{Oettel et~al.}(2007)\citenamefont{Oettel,
2596: Dom{\'\i}nguez, Tasinkevych, and Dietrich}}]{ODTD07}
2597: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
2598: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
2599: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tasinkevych}},
2600: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2601: \bibinfo{journal}{preprint} (\bibinfo{year}{2007}).
2602:
2603: \bibitem[{\citenamefont{Danov et~al.}(2006)\citenamefont{Danov, Kralchevsky,
2604: and Boneva}}]{DKB06}
2605: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}},
2606: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
2607: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~P.}
2608: \bibnamefont{Boneva}}, \bibinfo{journal}{Langmuir}
2609: \textbf{\bibinfo{volume}{22}}, \bibinfo{pages}{2653} (\bibinfo{year}{2006}).
2610:
2611: \bibitem[{\citenamefont{Frydel et~al.}(2007)\citenamefont{Frydel, Dietrich and Oettel}}]{FDO07}
2612: \bibinfo{author}{\bibfnamefont{D.} \bibnamefont{Frydel}},
2613: \bibinfo{author}{\bibfnamefont{S.} \bibnamefont{Dietrich}},
2614: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.}
2615: \bibnamefont{Oettel}}, \bibinfo{journal}{Phys.~Rev.~Lett.}
2616: \textbf{\bibinfo{volume}{99}}, \bibinfo{pages}{118302} (\bibinfo{year}{2007}).
2617:
2618: \bibitem[{\citenamefont{Dietrich}(1988)}]{Diet88}
2619: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
2620: \bibnamefont{in}
2621: \emph{\bibinfo{title}{Phase Transitions and Critical Phenomena}},
2622: \bibnamefont{edited by C.~Domb and J.~L.~Lebowitz}
2623: (\bibinfo{publisher}{Academic}, \bibinfo{address}{London},
2624: \bibinfo{year}{1988}),
2625: \bibnamefont{vol.}~\bibinfo{volume}{12},
2626: \bibnamefont{p.}~\bibinfo{pages}{1}.
2627: \end{thebibliography}
2628:
2629: \end{document}
2630:
2631: