cond-mat0611329/ms.tex
1: 
2: \documentclass[pre,nofootinbib,superscriptaddress,showpacs,showkeys]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5: 
6: 
7: %\newcommand{\rcite}[1]{Ref.~\citenum{#1}}
8: \newcommand{\rcite}[1]{Ref.~\onlinecite{#1}}
9: \newcommand{\rcites}[1]{Refs.~\onlinecite{#1}}
10: \newcommand{\eq}[1]{Eq.~(\ref{#1})}
11: 
12: %\newcommand{\prl}{{Phys.\ Rev.\ Lett.}}
13: %\newcommand{\pre}{{Phys.\ Rev.\ E}}
14: \newcommand{\epl}{{EPL}}
15: 
16: \newcommand{\br}{{\bf r}}
17: \newcommand{\bF}{{\bf F}}
18: \newcommand{\brp}{{\bf r}_\parallel}
19: \newcommand{\en}{{\bf e}_n}
20: \newcommand{\et}{{\bf e}_t}
21: \newcommand{\ez}{{\bf e}_z}
22: \newcommand{\ex}{{\bf e}_x}
23: \newcommand{\ey}{{\bf e}_y}
24: \newcommand{\er}{{\bf e}_r}
25: \newcommand{\bt}{{\bf t}}
26: \newcommand{\bn}{{\bf n}}
27: \newcommand{\nablap}{\nabla_\parallel}
28: \newcommand{\stress}{\mathsf{T}}
29: %\newcommand{\stress}{\stackrel{\leftrightarrow}{\bf T}}
30: \newcommand{\rdrop}{R_\mathrm{d}}
31: \newcommand{\vepsf}{\boldsymbol{\varepsilon}_\mathrm{el}}
32: \newcommand{\vepse}{\boldsymbol{\varepsilon}_\mathrm{ext}}
33: \newcommand{\vepsp}{\boldsymbol{\varepsilon}_\Pi}
34: \newcommand{\upd}{{\rm d}}
35: 
36: \begin{document}
37: 
38: \title{Force balance of particles trapped at fluid interfaces}
39: 
40: \author{Alvaro Dom{\'\i}nguez}
41: \email{dominguez@us.es}
42: \affiliation{F\'\i sica Te\'orica, Universidad de Sevilla, Apdo.~1065, 
43:   E--41080 Sevilla, Spain}
44: \author{Martin Oettel}
45: \affiliation{Institut f\"ur Physik, Universit{\"a}t Mainz, 
46:   D-55029 Mainz, Germany}
47: \author{S.\ Dietrich}
48: \affiliation{Max--Planck--Institut f\"ur Metallforschung, 
49:   Heisenbergstr.\ 3, D-70569 Stuttgart, Germany}
50: \affiliation{Institut f\"ur Theoretische und Angewandte Physik, 
51:   Universit{\"a}t Stuttgart, Pfaffenwaldring 57,
52:   D--70569 Stuttgart, Germany}
53: 
54: \date{Mar 21, 2008}
55: 
56: \pacs{82.70.Dd; 68.03.Cd}
57: \keywords{Colloids; surface tension and related phenomena}
58: 
59: \begin{abstract}
60:   We study the effective forces acting between colloidal particles trapped at a fluid
61:   interface which itself is exposed to a pressure field.
62:   To this end we apply what we call the ``force approach'', which relies solely on
63:   the condition of mechanical equilibrium and turns to be in a certain sense
64:   less restrictive than the more frequently used ``energy approach'',
65:   which is based on the minimization of a free energy functional.
66:   The goals are (i) to elucidate the advantages and disadvantages of the force
67:   approach as compared to the energy approach, and (ii) to disentangle
68:   which features of the interfacial deformation and of the
69:   capillary--induced forces between the particles follow from
70:   the gross feature of
71:   mechanical equilibrium alone, as opposed to features which depend on details
72:   of, e.g., the interaction of the interface with the
73:   particles or the boundaries of the system.
74:   First, we derive a general stress--tensor formulation of the forces
75:   at the interface. On that basis we work out a useful analogy with 2D 
76:   electrostatics in the particular case of small deformations of the interface 
77:   relative to its flat configuration. We
78:   apply this analogy in order to compute the asymptotic decay of
79:   the effective force between particles
80:   trapped at a fluid interface,
81:   extending the validity of previous results
82:   and revealing the advantages and limitations of the force approach 
83:   compared to the energy approach.
84:   It follows the application of the force approach to the case of
85:   deformations of a non--flat interface.
86:   In this context we first compute
87:   the deformation of a spherical droplet due to the
88:   electric field of a charged particle trapped at its surface and conclude
89:   that the interparticle capillary force is unlikely to explain certain recent
90:   experimental observations within such a configuration.
91:   We finally discuss the application of our approach to a generally curved
92:   interface and show as an illustrative example that a nonspherical
93:   particle deposited on an interface forming a minimal surface is pulled to 
94:   regions of larger curvature.
95: \end{abstract}
96: 
97: 
98: \maketitle 
99: 
100: 
101: \section{Introduction}
102: \label{sec:intro}
103: 
104: Experimental evidence has been accumulated that electrically charged,
105: $\mu$m--sized colloidal particles trapped at fluid interfaces can
106: exhibit long--ranged attraction despite their like charges
107: \cite{GhEa97,RGI97,SDJ00,QMMH01,NBHD02,TAKK03,FMMH04,CTNF05}. The
108: mechanisms leading to this attraction are not yet fully understood.
109: An attraction mediated by the interface deformation was proposed
110: \cite{NBHD02},
111: in analogy to the capillary force due to the weight of large floating
112: particles \cite{Nico49,CHW81}. However, for the particles sizes used
113: in the aforementioned experiments gravity is irrelevant. Instead,
114: one is led to invoke
115: electrostatic forces which act on the interface. This feature has
116: triggered investigations of capillary deformation and
117: capillary--induced forces beyond the well studied case of an
118: interface simply under the effects of gravity and surface tension
119: \cite{MeAi03,FoWu04,ODD05a,ODD05b,WuFo05,DOD05,ODD06,DKB06,DOD06a}.
120: These studies have relied almost exclusively on what we shall call the {\em
121:   ``energy approach''}, which is based on the minimization of a free
122: energy functional obtained as a parametric function of the positions
123: of the particles by integrating out the interfacial degrees of
124: freedom, leading to a ``potential of mean force''.
125: This functional has to include the contribution by the
126: interface itself, by the particles, and by the boundaries of the system.
127: Moreover, due to technical challenges the theoretical implementation of this approach 
128: is de facto restricted
129: to the regime of small interfacial deformations.
130: 
131: In the following, as an alternative we investigate
132: the {\em ``force approach''} which follows by directly applying the
133: condition of mechanical equilibrium.
134: Our analysis is based on the pressure field $\Pi(\br)$ (generated, e.g., by
135: electrostatic forces) acting on the interface between two fluid
136: phases. In general, the condition that an arbitrary piece ${\cal S}$
137: of this interface is in mechanical equilibrium reads (see, e.g., \rcite{Sege77})
138: \begin{equation}
139:   \label{eq:equilibrium}
140:   \int_{\cal S} \upd A \; \en \Pi + 
141:   \gamma \oint_{\partial{\cal S}} \upd\ell \; \et \times \en = 
142:   {\bf 0} ,
143: \end{equation}
144: where $\en$ is the local unit vector normal to the interface, $\et$ is the
145: unit vector tangent to the boundary $\partial {\cal S}$ (oriented such
146: that $\et \times \en$ points towards the exterior of ${\cal S}$),
147: $\upd A$ is the element of the interfacial area, $\upd\ell$ is the element
148: of the arclength along the contour $\partial {\cal S}$, and $\gamma$ is
149: the (spatially homogeneous) surface tension of the interface. 
150: In \eq{eq:equilibrium}, the first term is the so-called {\em bulk}
151: force exerted on the piece ${\cal S}$ by the pressure $\Pi$ and the
152: second one is the {\em line} force exerted on the contour and
153: generated by the surface tension (also called capillary force).
154: This equation is the starting point for the subsequent calculations.
155: 
156: The force approach allows us to obtain new results, to derive
157: previous ones more easily than within the energy approach,
158: and to gain additional insight. 
159: This approach was employed in \rcite{KPDI93} for the special case that gravity is
160: the only relevant force and it was shown to give the same results as
161: the energy approach if the deformations with respect to a flat interface are
162: small everywhere. For an arbitrary pressure field acting on the interface, in
163: \rcite{DOD05} we applied the force approach in order to obtain the deformation
164: of an otherwise flat interface far from the particles generating it.
165: In the following we further illustrate the force approach:
166: In Sec.~\ref{sec:stress} we first express the force exerted by the
167: interface in terms of a stress--tensor formulation, extending a recent
168: result \cite{MDG05a} to the most general case $\Pi(\br)
169: \neq$~constant. This formulation also allows us to establish a useful
170: analogy between two--dimensional electrostatics and the description of
171: small capillary deformations of a flat interface. Since this analogy has been
172: already employed by several authors in a more or less explicit manner,
173: here we present a thorough discussion addressing not only the issue of
174: interfacial deformations, but also that of boundary conditions and of the
175: capillary forces.
176: In Sec.~\ref{sec:force} we exploit the stress--tensor formulation and the electrostatic
177: analogy in order to study the interface--mediated effective force between
178: colloidal particles trapped at a fluid interface and to provide a 
179: detailed comparison with
180: the corresponding results obtained within the energy approach. 
181: In Sec.~\ref{sec:nonflat} we compute
182: the deformation of a spherical droplet due to the presence of a
183: charged particle at its surface, generalizing a corresponding result obtained in
184: \rcite{DOD05} for a flat interface and correcting certain claims in
185: the literature. Finally we discuss the more general case that the
186: unperturbed interface is curved.
187: Sec.~\ref{sec:end} provides a summary and an outlook.
188: 
189: 
190: \section{Stress--tensor formulation and electrostatic analogy}
191: \label{sec:stress}
192: 
193: The capillary force exerted by the interface (second term in
194: \eq{eq:equilibrium}) can be rewritten as
195: \begin{equation}
196:   \label{eq:defT}
197:   \gamma \oint_{\partial{\cal S}} \upd\ell \; \et \times \en = 
198:   \mbox{} - \oint_{\partial{\cal S}} 
199:   \upd\ell \; (\et \times \en) \cdot \stress , %= {\bf 0} ,
200: \end{equation}
201: which serves to define the stress tensor $\stress (\br) := \mbox{} -
202: \gamma \mathsf{1}(\br)$, where $\mathsf{1}(\br \in {\cal S})$ is the
203: 2D identity tensor on the tangent plane of ${\cal S}$ at each point
204: $\br$.
205: In these terms, the condition of mechanical equilibrium (\eq{eq:equilibrium})
206: takes the form
207: \begin{equation}
208:   \label{eq:equilibrium2}
209:   \int_{\cal S} \upd A \; \en \Pi =
210:   \oint_{\partial{\cal S}} 
211:   \upd\ell \; (\et \times \en) \cdot \stress . 
212: \end{equation}
213: We recall that $\et \times \en$ is a vector tangent to the surface
214: ${\cal S}$ but normal to the contour $\partial{\cal S}$ and pointing
215: outwards. This allows one to reinterpret \eq{eq:equilibrium2} as the definition of the stress tensor $\stress$, the flux of which
216: through a closed boundary is the bulk force (first term
217: in \eq{eq:equilibrium}) acting on the piece of interface enclosed by
218: that boundary.
219: In dyadic notation one has (summation over repeated indices is
220: implied)
221: \begin{equation}
222:   \stress =  \mbox{} - \gamma \mathsf{1} = 
223:   \mbox{} - \gamma \, g^{ab} \, {\bf e}_a \, {\bf e}_b 
224:   \quad \Rightarrow \quad
225:   T^{ab} =  \mbox{} - \gamma \, g^{ab} ,
226: \end{equation}
227: where $\{{\bf e}_1, {\bf e}_2\}$ is a local basis, at each
228: point tangent to the surface ${\cal S}$, and $g_{ab} = {\bf e}_a \cdot {\bf
229:   e}_b$ is the induced metric
230: ($g^{ab}$ are the contravariant components of this tensor).
231: In this form, we have the same stress
232: tensor as the one derived in \rcite{MDG05a} using methods of
233: differential geometry within the energy approach and for the
234: restricted case $\Pi(\br)=\rm{constant}$. 
235: We remark, however, that \eq{eq:equilibrium2} holds for an arbitrary
236: pressure field $\Pi(\br)$.
237: 
238: An analogy with 2D electrostatics emerges by considering 
239: small deformations relative to a {\it flat} interface\footnote{More
240:   precisely, the small quantity is the spatial gradient of the
241:   deformation (see Eqs.~(\ref{eq:quasiflat}, \ref{eq:Eanalogy})).}
242: corresponding to the generic experimental set-up, 
243: i.e., a situation like the one in Fig.~\ref{fig:ref} but with, e.g., a charged
244: particle \cite{GhEa97,RGI97,ACNP00,QMMH01,TAKK03,FMMH04,CTNF05}, a
245: nonspherical particle \cite{SDJ00,LYP06}, or a droplet at a nematic interface \cite{SCLN04}, so that the interface is
246: deformed by an electric field, by a nonplanar contact line,
247: or by the elastic stress in the nematic phase,
248: respectively. To this end, we identify the flat interface with the
249: $XY$--plane, so that any point of the deformed interface can be expressed as $\br
250: =\brp + \ez u(\brp)$ with $\brp := x\ex + y\ey$, and
251: (the subscript $_\parallel$ will be used to denote quantities
252: evaluated at and operators acting in the reference, flat interface)
253: \begin{subequations}
254:   \label{eq:quasiflat}
255: \begin{eqnarray}
256:   \en \, \upd A & = & 
257:   \left(\frac{\partial \br}{\partial x} \times 
258:     \frac{\partial \br}{\partial y} \right) 
259:   \upd y \upd x = 
260:   \left[ \ez - \nablap u \right] \upd A_\parallel , \\
261:   & & \nonumber \\
262:   \et \, \upd\ell & = & 
263:   (\upd\boldsymbol{\ell}_\parallel \cdot \nabla_\parallel) \br = 
264:   \upd\boldsymbol{\ell}_\parallel + 
265:   \ez (\upd\boldsymbol{\ell}_\parallel \cdot \nablap u) ,
266: \end{eqnarray}
267: \end{subequations}
268: where $\upd \boldsymbol{\ell}_\parallel = \ex \upd x + \ey \upd y$,
269: $\upd A_\parallel = \upd x \upd y$, and $\nablap = \ex
270: (\partial/\partial x) + \ey (\partial/\partial y)$. 
271: We denote the projection of any piece of interface ${\cal S}$ onto the
272: $XY$--plane as ${\cal S}_\parallel$, and introduce the unit vector
273: $\bn$ in the $XY$--plane which is normal to the contour $\partial{\cal
274:   S}_\parallel$ and points outwards. With this, we expand
275: \eq{eq:equilibrium2} in terms of the deformation $u(\brp)$; to lowest order 
276: the component in the direction of $\ez$ is linear in $u(\brp)$,
277: \begin{subequations}
278:   \label{eq:Eanalogy}
279:   \begin{equation}
280:     \label{eq:force_z}
281:     \int_{{\cal S}_\parallel} \upd A_\parallel \; \Pi =
282:     \mbox{} - \gamma \oint_{\partial{\cal S}_\parallel} 
283:     \upd\ell_\parallel \; \bn \cdot \nablap u .
284: \end{equation}
285: The local version of this equality is the linearized
286: Young--Laplace equation:
287: \begin{equation}
288:   \label{eq:linYL}
289:   \gamma \nablap^2 u = - \Pi .   
290: \end{equation}
291: To lowest order the components of \eq{eq:equilibrium2} in the
292: $XY$--plane are quadratic in the deformation:
293: \begin{equation}
294:   \label{eq:force_xy}
295:   \mbox{} - 
296:   \int_{{\cal S}_\parallel} \upd A_\parallel \; \Pi \; \nablap u = 
297:   \oint_{\partial{\cal S}_\parallel} 
298:   \upd\ell_\parallel \; \bn \cdot \stress_{\parallel} , 
299: \end{equation}
300: where
301: \begin{equation}
302:   \label{eq:Tpar}
303:   \stress_{\parallel}\ := \gamma \left[ 
304:     (\nablap u) (\nablap u) - \frac{1}{2} |\nablap u|^2 
305:     \; \mathsf{1}_\parallel \right] 
306: \end{equation}
307: \end{subequations}
308: is a stress tensor defined in the $XY$--plane. We remark that
309: \eq{eq:force_xy} also implies \eq{eq:linYL} upon applying Gauss'
310: theorem, demonstrating consistency.
311: 
312: The form of \eq{eq:linYL} allows us to identify $u$ with an
313: electrostatic potential 
314: (and $-\nablap u$ with an electric field), 
315: $\Pi$ with a charge density 
316: (``capillary charge'' \cite{KDD01}), and $\gamma$ with a 
317: permittivity\footnote{There is also a dual magnetostatic
318:   interpretation in terms of magnetic fields created by currents along
319:   the $Z$--direction; the correspondences are $\ez u(\brp)
320:   \leftrightarrow {\bf A}(\brp)$, $\nablap \times (\ez u)
321:   \leftrightarrow {\bf B}(\brp)$, $\gamma \leftrightarrow 1/\mu_0$,
322:   $\ez \Pi(\brp) \leftrightarrow {\bf j}(\brp)$.}. The boundary
323: conditions usually imposed on
324: $u$ at a contour ${\cal C}_\parallel$ have a close electrostatic
325: analogy, too (see Fig.~\ref{fig:bc} for the notation):
326: 
327: \begin{figure}
328:   \centering{\includegraphics[width=.85\textwidth]{figure1.eps}}
329:   \caption{Generic view of the three-phase contact line of an
330:     interface with a solid boundary. (a) Top view: projection ${\cal
331:       S}_\parallel$ onto the $XY$-plane of the interface (dashed area)
332:     near the projected contact line ${\cal C}_\parallel$ (dashed
333:     line). The unit vector $\bn$ is normal to the projected contact
334:     line and is directed towards the exterior of ${\cal S}_\parallel$.
335:     (b) Side view of the same configuration within a vertical plane
336:     containing $\bn$. The thin horizontal line is the section of
337:     ${\cal S}_\parallel$ (i.e., the flat, unperturbed interface),
338:     while the dash--dotted line is the section of the actual interface
339:     ${\cal S}$, which can be approximated locally by the tangent
340:     forming an angle $\theta$ (contact angle) with the solid boundary.
341:     The latter is the full thick line, which looks locally like a
342:     straight line, in general inclined by an angle $\omega$ with
343:     respect to the vertical direction (dotted line). The height of the interface
344:     at contact entering into the boundary
345:     condition~(\ref{eq:pinnedbc}) is $z_\parallel:=z(\brp)$.  The
346:     boundary condition~(\ref{eq:anglebc}) expresses the slope of the
347:     interface at contact with $\alpha := \theta + \omega$.}
348:   \label{fig:bc}
349: \end{figure}
350: 
351: 
352: \noindent
353: (i) The potential is given at ${\cal C}_\parallel$. $\Leftrightarrow$
354: The interface is pinned at the contour at a height $z(\brp)$:
355: \begin{equation}
356:   \label{eq:pinnedbc}
357:   u(\brp) = z (\brp), \qquad \brp \in {\cal C}_\parallel .
358: \end{equation}
359: 
360: \noindent
361: (ii) The normal component of the electric field is given at ${\cal
362:   C}_\parallel$.  $\Leftrightarrow$ The contact angle $\theta(\brp)$
363: is specified at each point of the contour: 
364: \begin{equation}
365:   \label{eq:anglebc}
366:   (\bn \cdot \nablap) u(\brp) = \cot \alpha(\brp), \qquad 
367:   \brp \in {\cal C}_\parallel ,
368: \end{equation}
369: where the angle $\alpha(\brp)$, defined in Fig.~\ref{fig:bc}, must be
370: close to $\pi/2$ for reasons of consistency with the approximation of
371: small deformations.
372: 
373: This means that the equivalence with the corresponding electrostatic
374: analogy is exact concerning the relationship between the deformation
375: field $u(\brp)$ and its sources (i.e., the pressure field $\Pi$ and
376: the boundary conditions).
377: The analogy carries over almost exactly, too, to the elastic forces in the $XY$--plane
378: (``lateral capillary forces'') arising from interfacial deformation:
379: according to Eqs.~(\ref{eq:force_xy},
380: \ref{eq:Tpar}), $\stress_\parallel$ corresponds to Maxwell's stress
381: tensor, the flux of which through a closed boundary gives the electric
382: force acting on the enclosed charge. However, the force related to the
383: deformation is actually the interfacial stress, which is {\rm minus}
384: the flux of this tensor (see \eq{eq:defT}). Therefore, the
385: electrostatic analogy is valid up to a reversal of the forces and the
386: peculiarity arises that ``capillary charges'' will attract~(repel)
387: each other if they have equal~(different) sign. The origin of this
388: peculiarity is that \eq{eq:equilibrium2}, which  in
389: the spirit of electrostatics can be reinterpreted as a definition of $\stress$, 
390: is actually
391: a relationship between bulk and capillary forces as two physically different 
392: kinds of forces. The actual connection of $\stress$ with a
393: force is \eq{eq:defT}.
394: 
395: We note that the electrostatic analogy holds wherever the
396: deformation of the interface is small, even if there are other regions
397: of the interface where this is not true. Such ``nonlinear patches''
398: can be surrounded by contours where the deformation is small, so that
399: the values of the field $u$ and of its derivatives at these contours
400: play the role of a boundary condition for the ``linear patches''.
401: This means that the ``nonlinear patches'' are replaced by a
402: distribution of virtual ``capillary charges'' inside the corresponding
403: regions. In particular, there is a simple physical meaning associated
404: with the total ``capillary charge'' and the total ``capillary
405: dipole''. The ``capillary charge'' $Q$ of a piece of interface bounded
406: by a contour ${\cal C}$ is given by Gauss' theorem solely in terms of
407: the value of the deformation at the contour (see \eq{eq:force_z}):
408: \begin{equation}
409:   \label{eq:gauss}
410:   Q = \mbox{} - \gamma \oint_{{\cal C}_{\parallel}} 
411:   \upd\ell_\parallel \; \bn \cdot \nablap u .
412: \end{equation}
413: The right hand side of this equation is minus the capillary force
414: exerted on the piece of interface in the $Z$--direction. This implies
415: that in terms of the bulk force ${\bf F}_{\rm bulk}$ and by virtue of
416: the condition of mechanical equilibrium one has
417: \begin{equation}
418:   \label{eq:mono_general}
419:   Q = \ez \cdot {\bf F}_{\rm bulk} .
420: \end{equation}
421: This holds even if the deformation in the bulk (i.e., inside the
422: contour ${\cal C}$) is not small. In the same manner, it can be shown
423: (see Appendix~\ref{app:dipole})
424: that the total ``capillary dipole'' ${\bf P}$ (with respect to the
425: origin of the coordinate system) and the torque ${\bf M}_{\rm bulk}$
426: (with respect to the same origin) exerted by
427: the bulk force are related according to
428: \begin{equation}
429:   \label{eq:dipole}
430:   {\bf P} = \ez \times {\bf M}_{\rm bulk} .
431: \end{equation}
432: 
433: The electrostatic analogy provides a transparent visualization of
434: small interfacial deformations and ensuing forces in terms of a 2D
435: electrostatic problem. We note that in \rcite{Paun98} such a kind of
436: analogy is also established, but between the capillary interaction and
437: the 3D electrostatic problem (DLVO theory, see, e.g., \rcite{KlLa03})
438: for that very same 3D geometrical set-up.
439: As a side remark we note that if gravity (or the disjoining pressure due to a substrate)
440: is relevant, it contributes a pressure field which depends explicitly
441: on the deformation field: $\Pi_\mathrm{grav}= - \gamma u / \lambda^2$
442: with the capillary length $\lambda$ \cite{DOD06a}. This replaces
443: \eq{eq:linYL} by a different equation which is formally equivalent to
444: the field equation of the Debye--H\"uckel theory for dilute
445: electrolytes (see, e.g., \rcite{KlLa03}) with $\lambda$ playing the role of the Debye length. This
446: suggests that extending the electrostatic analogy to this case
447: is possible, but this task is beyond the scope of the present effort.
448: 
449: \section{Effective interparticle forces}
450: \label{sec:force}
451:  
452: \begin{figure}
453:   \centering{\includegraphics[width=.65\textwidth]{figure2.eps}}
454:   \caption{Configuration of two particles (top view) a lateral
455:     distance $d$ apart, trapped at an asymptotically flat interface.
456:     The left half ${\cal S}$ of the full interface is bounded by the
457:     particle--interface contact line ${\cal C}_0$ (generically
458:     noncircular), the oriented projection of which onto the
459:     $XY$--plane is ${\cal C}_{0,\parallel}$, and a contour ${\cal C}$,
460:     the oriented projection ${\cal C}_\parallel$ of which consists of
461:     a piece of the $Y$--axis and a circle of radius $L\to\infty$. The
462:     unit vector $\bn$ is normal to the contour and directed towards
463:     the exterior of ${\cal S}$. The origin of the coordinate system is
464:     located at the midpoint $O$, while the coordinates $r_R$,
465:     $\varphi_R$ parametrize the plane with respect the center of the
466:     right particle.}
467:   \label{fig:2coll}
468: \end{figure}
469: 
470: The capillary deformation may give rise to an effective attraction
471: between two identical particles trapped at the interface which could
472: explain certain corresponding experimental observations. We compute this force using
473: the 
474: electrostatic analogy derived
475: above. The study of this issue will also clarify
476: the general relationship between the force and the energy approaches as well as
477: the respective advantages and disadvantages.
478: We consider the equilibrium configuration of two 
479: particles at
480: an asymptotically flat interface and fixed to be a distance $d$ apart (see Fig.~\ref{fig:2coll}).
481: The total capillary force $\bF_\mathrm{total}$ acting on the left half (which includes the piece
482: ${\cal S}$ {\em and} the enclosed particle) is given by \eq{eq:defT}:
483: \begin{equation}
484:   \label{eq:Ftotal}
485:   \bF_\mathrm{total} = 
486:   \mbox{} - 
487:   \oint_{{\cal C}} 
488:   \upd\ell \; (\et \times \en) \cdot \stress .
489: \end{equation}
490: If the particle separation $d$ is large enough, we can assume that the
491: deformation of the interface {\em at} (but not necessarily inside) the
492: contour ${\cal C}$
493: is small, so that the lateral force
494: $\bF_\parallel$ (= component of $\bF_\mathrm{total}$ in the
495: $XY$--plane = $\bF_\mathrm{total}-(\bF_\mathrm{total}\cdot\ez)\ez$) 
496: is (see \eq{eq:force_xy})
497: \begin{equation}
498:   \label{eq:Flateral}
499:   \bF_\parallel \approx  \mbox{} - \oint_{{\cal C}_\parallel} 
500:   \upd\ell_\parallel \; \bn \cdot \stress_\parallel . 
501: \end{equation}
502: In the limit $L\to\infty$, the contribution from the circular part of ${\cal C}_\parallel$ vanishes and only the knowledge of the deformations at the 
503: straight midline part of ${\cal C}_\parallel$ is required, for
504: which the electrostatic analogy will hold provided the interfacial
505: deformations are small there.
506: Thus, one can try to estimate the ``electric potential'' $u$ at the
507: midline as $d\to\infty$
508: by a multipole expansion. 
509: The deformation field $u_R$ created by the right
510: half plane 
511: behaves asymptotically like (see Appendix~\ref{app:multipolar})
512: \begin{equation}
513:   \label{eq:multipole}
514:   u_R(\br_R) \sim \frac{Q_0}{2\pi\gamma} \ln \frac{\zeta}{r_R}
515:   + \frac{1}{2\pi\gamma} \sum_{s=1}^\nu \frac{Q_s \,{\rm e}^{-i s \varphi_R} + Q_s^* \,{\rm e}^{i s \varphi_R}}{2 s (r_R)^s} 
516:   + \Delta u_R .
517: \end{equation}
518: Here, ${\bf r}_R = \brp - \ex d/2$ is the position of a point relative
519: to the center of the right particle (see Fig.~\ref{fig:2coll}), $\zeta$ is a
520: fixed length determined by the distant boundary conditions, which set the
521: zero point (undeformed interface) of the ``electric potential'' $u_R$, and $Q_s$ is given in terms
522: of the $2^s$--pole of ``capillary
523: charge'' associated with the particle and the surrounding interfacial deformation; if this is small everywhere, it is
524: \begin{equation}
525:   \label{eq:charge}
526:    Q_s = \int \upd A_R \; 
527:    (r_R)^s \, {\rm e}^{i s \varphi_R} \, \Pi_{\rm (single)} (\br_R) 
528:    + \tilde{Q}_s^{\rm (single)},
529: \end{equation}
530: with $\tilde{Q}_s^{\rm (single)}$ the corresponding charge associated to the
531: particle (which is defined by the multipole
532: expansion (see \eq{eq:multexpansion}) applied to the interfacial
533: deformation at the contact line).
534: The multipole expansion is based on the implicit assumption that the main
535: source of the deformation field is concentrated at or near the
536: particles. For that reason $Q_s$ is given by the ``capillary charge''
537: distribution for $d=\infty$, i.e., for the single--particle
538: configuration.
539: The asymptotically subdominant term $\Delta u_R$ in \eq{eq:multipole}
540: accounts for the corrections to this assumption, i.e.,
541: ``polarization'' effects by the second particle and the fact that,
542: even in the single--particle configuration,
543: the pressure field $\Pi$ is expected to decay smoothly 
544: asymptotically far from the particles rather
545: than dropping exactly to zero beyond some distance: if $\Pi
546: (r_\parallel\to\infty) \sim r_\parallel^{-n}$ (actually $n=6$ in
547: realistic models in the case of electric stresses
548: \cite{Hurd85,ACNP00,DaKr06a,DFO08}, and $n=8$ in the case of nematic stresses
549: \cite{ODTD07}), there is the bound $\nu<n-2$ 
550: (see the sum in \eq{eq:multipole} and Appendix~\ref{app:multipolar}).
551: 
552: In general the leading term in the multipole expansion is determined by the
553: ``capillary monopole'' $Q_0$. 
554: Thus asymptotically for $d\to\infty$, $\bF_\parallel$ will be given by
555: the ``electric force'' between two monopoles:
556: \begin{equation}
557:   \label{eq:Flog}
558:   \bF_\parallel \approx \frac{Q_0^2}{2\pi\gamma d} \ex ,
559: \end{equation}
560: with the reversed sign as explained in the previous Section, so that
561: it describes an attraction. Since $Q_0$ is given in terms of the net
562: bulk force according to \eq{eq:mono_general}, the $1/d$-dependence
563: only arises if there is an external field acting on the system. For
564: example, \eq{eq:Flog} corresponds to the ``flotation force'' (at
565: separations $d$ much smaller than the capillary length) for which $\bF_{\rm
566:   bulk}$ is due to gravity.
567: On the contrary, $Q_0$ will vanish if the system is mechanically
568: isolated (so that any bulk forces acting on the particle are, according to the
569: action--reaction principle, equal --- but of oppositte sign --- to any bulk
570: forces acting on the interface and hence $\bF_{\rm bulk}={\bf 0}$).
571: This was the key point of a recent controversy as the force in
572: \eq{eq:Flog} was advocated to explain certain experimental
573: observations \cite{NBHD02,DKB04,SCLN04} while missing that mechanical
574: isolation (as purported in the experiments) rules out this force
575: \cite{MeAi03,FoWu04,ODD05a,ODD06,ODTD07}.
576: 
577: If $Q_0=0$, the capillary force is determined by higher-order terms in
578: the multipole expansion~(\ref{eq:multipole}). Mechanical isolation
579: implies the vanishing of the net bulk force and torque, i.e., ``capillary
580: monopole'' and ``capillary dipole'' (see Eqs.~(\ref{eq:mono_general},
581: \ref{eq:dipole})). Thus in general, $\bF_\parallel$ will take the form of
582: a force between quadrupoles, i.e., it is anisotropic and scales as
583: \begin{equation}
584:   \label{eq:Fquad}
585:   |\bF_\parallel| \propto \frac{|Q_2|^2}{d^5}.
586: \end{equation}
587: An experimental realization of this case corresponds to nonspherical
588: inert particles, so that $\Pi\equiv 0$ and the
589: interfacial deformation is solely due to an undulated contact line (for
590: a recent corresponding experimental study see \rcite{LYP06}). Concerning
591: the possibility to relate corresponding experimental observations
592: and theoretical descriptions we point out the
593: difficulty that $Q_2$ and higher ``capillary poles'' are
594: known only in terms of the deformation field $u$, in contrast to $Q_0$
595: and $Q_1$, which are given by the directly measurable and independently accessible 
596: quantities ``bulk force'' and ``bulk torque'', respectively.
597: 
598: Another experimentally relevant situation of mechanical isolation
599: corresponds to the case that the ``capillary charges'' $Q_s$ of all
600: orders $s\geq 0$ vanish. For example, for an electrically charged,
601: spherical particle in mechanical isolation one has $Q_0=0$ and $Q_1=0$ by mechanical isolation, 
602: and by symmetry a rotationally invariant interfacial deformation in
603: the single--particle configuration, giving $Q_s=0$ also for any
604: $s\geq 2$. In this case \eq{eq:multipole} reduces to the correction
605: $\Delta u_R$ and the computation of $\bF_\parallel$ requires a specific model for
606: $\Pi(\brp)$ and a detailed calculation. In view of our present purposes, we
607: qualitatively derive only a bound on how rapidly $\bF_\parallel$
608: decays as function of the separation $d$.
609: 
610: The correction $\Delta u_R(r_R\to\infty)$ has a contribution $\sim
611: (r_R)^{2-n}$ already in the single--particle configuration because the
612: rotationally symmetric ``charge density'' $\Pi(r_R\to\infty) \sim
613: (r_R)^{-n}$ does not have a compact support (see
614: Appendix~\ref{app:multipolar}).  This provides a contribution to the
615: capillary force $\bF_\parallel$ which is equal to the force between the
616: ``capillary charges'' at each side of but near the midline because 
617: the presence of the second particle breaks the rotational symmetry.
618: This force can be estimated to decay like $d^{-1+2(2-n)}$
619: because the net ``capillary charge'' in the region farther than a
620: distance $d$ from one particle is $\sim \int_{r_\parallel > d}
621: dA_\parallel \; \Pi \sim d^{2-n}$ and the force between charges decays
622: $\sim d^{\mbox{}-1}$.
623: 
624: Additionally, $\Delta u_R$ has genuine two--particle contributions
625: ``induced'' by the second particle which can be modeled by means of
626: ``induced capillary charges'' $Q_s(d)$ depending on the particle
627: separation. Generically the dominant term will be a ``capillary
628: dipole''\footnote{The net ``capillary monopole'' of the halfplane must
629:   vanish exactly due to mechanical isolation and symmetry reasons.
630:   Although the net torque on the whole system also vanishes, in this case
631:   symmetry considerations do not exclude that the net torque on
632:   one halfplane is opposite to the net torque on the opposite
633:   halfplane, so that a ``capillary dipole'' in one halfplane is
634:   possible.} giving rise to a correspondingly dipole-dipole force
635: $\sim |Q_1(d)|^2 /d^3$, where the ``induced dipole'' $Q_1(d)$ must
636: decay at least like the inducing field. This is caused by various reasons. 
637: If $Q_1(d)$ arises by a violation of the boundary conditions
638: at the contact line, one has $Q_1(d) = {\cal O}(\Delta u_R(d)) \sim
639: d^{2-n}$, and the force decays by a factor $1/d^2$ faster than the
640: contribution discussed above concerning the violation of rotational
641: symmetry.
642: But there can also be deviations from the linear superposition of the pressure on
643: the interface. This occurs, e.g., if the interfacial deformation is
644: due to electric fields emanating from the particles, so that $\Pi
645: \propto E^2$ and $\Pi(\br_\parallel) - [\Pi_{\rm (single)}(\br_R) +
646: \Pi_{\rm (single)}(\br_L)] \propto E_{\rm (single)}(\br_R) \, E_{\rm
647:   (single)}(\br_L)$. In this case one has $Q_1(d) = {\cal O}(E_{\rm
648:   (single)}(d)) \sim d^{-n/2}$.
649: Thus we can conclude that the lateral force must decay as function of separation
650: asymptotically at least as
651: \begin{equation}
652:   \label{eq:Fmu}
653:   F_\parallel = {\cal O}(d^{\mbox{}-\min{(2n-3,n+3)}}) ,
654: \end{equation}
655: depending on the value of $n$. In any case this force decays more
656: rapidly than the expression given by \eq{eq:Flog}.
657: 
658: It is instructive to compare $F_\parallel$ with the force obtained
659: within the energy approach. The latter consists of finding the parametric
660: dependence on $d$ of the free energy for the two--particle
661: configuration,
662: \begin{equation}
663:   \label{eq:Vmen}
664:   V_\mathrm{men} = \int_{{\cal S}_\parallel} \upd A_\parallel \; \left[ 
665:     \, \frac{\gamma}{2} \, |\nablap u|^2 - \Pi \, u \right] 
666:   + V_\mathrm{part} ,
667: \end{equation}
668: where the integral is the contribution by the interface and
669: $V_\mathrm{part}$ collects the direct contribution by the particles
670: \cite{ODD05a,DOD06a}. (Within the electrostatic analogy, the effect of
671: $V_\mathrm{part}$ would be replaced by appropriate boundary conditions
672: on the ``potential'' $u$ at the interface--particle contact lines.)
673: $V_\mathrm{men}(d)$ plays the role of a ``potential of mean force''
674: for the particle--particle interaction, giving rise to a corresponding ``mean
675: force''
676: \begin{equation}
677:   \label{eq:Fmen}
678:   F_\mathrm{men}(d) = \mbox{} - V'_\mathrm{men}(d) 
679: \end{equation}
680: upon integrating out the capillary degrees of freedom within thermal equilibrium. 
681: One should keep in mind that this approach captures only the mean--fieldlike contribution to the mean force. The capillary wavelike
682: fluctuations of the interface around the mean meniscus profile
683: generates additional, Casimir--like contributions to the force
684: \cite{LOD06,LeOe07}, which we do not consider in the following.
685: 
686: One can distinguish two cases. First, if there are ``permanent
687: capillary charges'' ($Q_0\neq 0$ if the system is not mechanically
688: isolated, or $Q_s\neq 0$ for some $s\geq 2$ for mechanical isolation,
689: as discussed above), $F_\parallel$ coincides with $F_\mathrm{men}$;
690: see, e.g., \rcites{Nico49,CHW81,ODD05a} for a derivation of
691: \eq{eq:Flog} or \rcites{SDJ00,KDD01,FoGa02,VSH05} for obtaining
692: \eq{eq:Fquad}
693: in the context of the energy approach with the simplifying assumption
694: that the interface deformation is small
695: everywhere\footnote{Reference~\onlinecite{KPDI93} performs an exhaustive comparison
696:   of the two approaches for the special case that gravity is the only
697:   source of deformation 
698:   and the interfacial deformation is small everywhere.}. The reasoning
699: presented here extends, however, this result also to the case that the
700: deformation around the particles is not small, requiring this only
701: near the midline between the particles, i.e., asymptotically for
702: $d\to\infty$.
703: Furthermore, the electrostatic analogy shows immediately 
704: that the capillary
705: forces $F_\parallel$ are asymptotically pairwise additive in
706: a configuration with more than two particles provided they
707: possess a nonvanishing ``permanent capillary pole''.
708: 
709: The second case corresponds to the absence of 
710: ``permanent capillary charges'' as described above. This has been thoroughly
711: investigated in \rcites{ODD05b,WuFo05,DOD06a} within the energy
712: approach, and has led to $F_\mathrm{men}(d) \sim d^{-1-n/2}$, which does
713: not agree with any of the possible asymptotic decays indicated in
714: \eq{eq:Fmu}.
715: In order to understand this discrepancy, we recall that by
716: definition (\eq{eq:Flateral}) $\bF_\parallel$ 
717: represents the net force acting on the subsystem formed by the
718: particle {\em and} the piece of interface enclosed by the contour
719: indicated in Fig.~\ref{fig:2coll}. The work done by this force upon an
720: infinitesimal virtual displacement $\delta d$ is not related in any
721: simple manner to the change $\delta V_\mathrm{men}$, which according
722: to the definition in \eq{eq:Vmen} will involve the work done by local
723: forces during the rearrangement of the ``capillary charges'' {\em
724:   inside} the subsystem, so that in general $F_\parallel \neq F_\mathrm{men}$. 
725: 
726: In conclusion one has $F_\parallel = F_\mathrm{men}$ if the $d$--dependence of
727: $V_\mathrm{men}(d)$ is dominated asymptotically by a multipole
728: expansion, i.e.,
729: the whole subsystem can be replaced by a set of point ``capillary
730: poles'': the degrees of freedom related to the internal
731: structure 
732: are irrelevant and only the separation $d$ and the orientation of the
733: ``capillary poles'' matter.
734: This is related to the validity of the ``superposition
735: approximation''\cite{Nico49} usually employed in the energy approach, which
736: consists of approximating the deformation field $u$ by the sum of the
737: deformation fields induced by each particle in the single--particle
738: configuration. In \rcite{DOD06a} it is shown that this
739: approximation is valid if the system is not mechanically isolated
740: because in that case, asymptotically for $d\to\infty$, the
741: interface--mediated effect of one particle on the other amounts to
742: shift it --- together with its surrounding interface ---
743: vertically as a whole, i.e., without probing or affecting the
744: ``internal structure'' of the subsystem ``particle plus surrounding
745: interface''.
746: 
747: Finally, it is clear that both $F_\parallel$, defined in
748: \eq{eq:Flateral}, and $F_\mathrm{men}$, defined in \eq{eq:Fmen},
749: include a contribution from $\Pi$
750: and thus differ from the force acting {\em only} on the colloidal
751: particle (which would be the integral in \eq{eq:Flateral} extended only along the
752: particle--interface contact line). 
753: If one is interested in physical situations in thermal equilibrium,
754: $F_\mathrm{men}$ does represent the effective force between the
755: particles, i.e., once the capillary degrees of freedom have been
756: integrated out. 
757: In dynamical situations out of equilibrium, completely new
758: considerations have to be made concerning, e.g., whether the capillary
759: degrees of freedom can be assumed to have relaxed towards thermal
760: equilibrium in the dynamical time scale of interest. But this
761: discussion lies beyond the scope of the present analysis.
762: 
763: 
764: 
765: \section{Nonplanar reference interface}
766: \label{sec:nonflat}
767: 
768: In the following we shall discuss some applications of \eq{eq:equilibrium}
769: for particles trapped at an interface which in its
770: unperturbed state is curved. 
771: In Subsec.~\ref{sec:droplet} we shall first consider the interfacial
772: deformation induced by a single charged particle on an otherwise
773: spherical droplet.
774: This configuration is particularly relevant for the experiment
775: described in \rcite{NBHD02}. As explained in the previous section,
776: mechanical isolation rules out a monopolarlike (i.e., logarithmic)
777: deformation if the unperturbed interface is flat. Since there has
778: been recently a controversy whether this conclusion is altered by the
779: curvature of the droplet, we shall present a thorough analysis for such systems.
780: In Subsec.~\ref{sec:curved} the application of the electrostatic
781: analogy to a generally curved interface is illustrated.
782: 
783: \subsection{Particle on a spherical droplet}
784: \label{sec:droplet}
785: 
786: \begin{figure}
787:   \centering{\includegraphics[width=.65\textwidth]{figure3.eps}}
788:   \caption{Charged colloidal particle (radius $R$) at the interface of
789:     a droplet residing on a plate. Positions on the interface are
790:     parametrized by the polar angle $\psi$ and the revolution angle
791:     $\varphi$.
792:     Without the colloidal particle and neglecting gravity
793:     the droplet has spherical shape
794:     with radius $\rdrop$ and normal vector $\er$. $\brp = \rdrop
795:     \er(\psi,\varphi)$ is a point at the unperturbed, spherical
796:     interface (dashed line) whereas $\br = \brp + u(\psi,\varphi) \er$ is a point at
797:     the perturbed, nonspherical interface (full line). The perturbed interface
798:     intersects the plate for $\psi=\psi_1$.}
799:     \label{fig:droplet}
800: \end{figure}
801: 
802: 
803: We consider a charged spherical 
804: particle of radius $R$ trapped at the interface of a droplet which
805: resides on a plate (Fig.~\ref{fig:droplet}). This configuration models
806: the experiment described in \rcite{NBHD02} in the absence of
807:   gravity.
808: Our goal is to compute the deformation of the droplet far from the
809: particle. Compared with the energy approach, the force approach has
810: two advantages:
811: (i) The result is more general because we have to assume only that the
812: deformation is small {\em far} from the particle; the usual linear
813: approximation is not required to hold also near the particle. (ii)
814: The boundary condition ``mechanical equilibrium of the particle'' is
815: incorporated easily 
816: irrespective of the details how the particle is attached to the
817: interface. It will turn out that the implementation of this condition
818: has been the source of mistakes in the literature.
819: 
820: We apply \eq{eq:equilibrium} to the piece ${\cal S}(\Psi)$ of the
821: curved interface bounded on one side by the particle--interface contact
822: line ${\cal C}_0$
823: and on the other side by a circle given by the constant latitude
824: $\Psi$, ${\cal C}(\Psi) := \{\psi=\Psi \leq \psi_1 \}$, so that
825: $\partial{\cal S} = {\cal C}_0 \cup {\cal C}(\Psi)$, and we assume that
826: the particle is located at the apex opposite to the plate
827: (Fig.~\ref{fig:droplet}).  The unperturbed state corresponds to an
828: uncharged particle which does not exert a force on the interface, so
829: that the equilibrium shape of the interface is spherical. In the
830: presence of electric charges, the interface will deform. If the
831: particle stays at the upper apex, also the deformed interface exhibits
832: axial symmetry.
833: We rewrite \eq{eq:equilibrium} in three steps. \\
834: 
835: \begin{figure}
836:   \centering{\includegraphics[width=.65\textwidth]{figure4.eps}}
837:   \caption{Side view of a single electrically neutral and spherical
838:     particle at a planar fluid interface in equilibrium. The radius
839:     $r_0 = R\sin\theta$ of the circular contact line follows from the
840:     equilibrium contact angle $\theta$. If the interface was that of
841:     a large droplet (radius $\rdrop\gg R$, see
842:     Fig.~\ref{fig:droplet}), this expression would exhibit correction
843:     terms of order $R/\rdrop$. In the same manner, the presence of
844:     charges would deform the interface and introduce corrections which
845:     in first order are proportional to the deformation.}
846:   \label{fig:ref}
847: \end{figure}
848: 
849: 
850: \noindent
851: (i) The pressure splits into 
852: \begin{equation}
853:   \label{eq:splitPi}
854:   \Pi (\br) = \Delta p + \Pi_\mathrm{el}(\br) , \qquad 
855:   \Delta p = \frac{2\gamma}{\rdrop} \left( 1- \mu/2\right) .
856: \end{equation}
857: Here, $\rdrop$ is the radius of the unperturbed, spherical droplet,
858: and $2\gamma/\rdrop$ is the pressure jump across the interface in the
859: unperturbed state. The dimensionless constant $\mu$ accounts for the
860: change in hydrostatic pressure due to enforcing the condition of
861: constant droplet volume in the presence of interface deformations.
862: $\Pi_\mathrm{el}(\br)$ is the pressure field created by the electric
863: field emanating from the particle which includes electric stresses and
864: an osmotic pressure due to a possible discontinuity of the ion
865: concentrations at the interface (see, e.g.,
866: \rcites{SMN02,Wuer06a}); this pressure field follows from solving the
867: corresponding electrostatic problem. We write
868: \begin{equation}
869:   \int_{{\cal S}} \upd A \; \en \Pi_\mathrm{el} = 
870:   2\pi\gamma r_0 \, \vepsp -
871:   \int_{\cal S_\mathrm{men}\backslash \cal S} \!\!\!\!\!\! \upd A \; 
872:   \en \Pi_\mathrm{el} ,
873: \end{equation}
874: where we have introduced the dimensionless electric force acting on
875: the whole interface ${\cal S}_\mathrm{men}$ (which at the apex has a
876: hole carved out),
877: \begin{equation}
878:   \label{eq:epspi}
879:   \vepsp := \frac{1}{2\pi\gamma r_0}
880:   \int_{\cal S_\mathrm{men}} \!\!\!\!\!\! \upd A \; 
881:   \en \Pi_\mathrm{el} .
882: \end{equation}
883: Here 
884: $r_0=R\sin\theta$,
885: where $\theta$ is the equilibrium contact angle between the fluid
886: phases and the particle; $r_0$ is actually the radius of the circular contact
887: line of the uncharged particle at a planar interface
888: (see Fig.~\ref{fig:ref}). \\
889: 
890: \noindent
891: (ii) Since the particle is in mechanical equilibrium, the
892: contact line force must be balanced by the hydrostatic force
893: $\bF_{\Delta p}$ and the electric force $\bF_\mathrm{el}$ acting on
894: the particle, as well as by any other force $\bF_\mathrm{ext}$ of
895: external origin (e.g., an optical tweezer pushing or pulling the particle):
896: \begin{equation}
897:   \label{eq:contact}
898:   \gamma \oint_{{\cal C}_0} \upd\ell \; \et \times \en = 
899:   \bF_{\Delta p} - 2\pi\gamma r_0 (\vepsf + \vepse) ,
900: \end{equation}
901: where we have introduced the dimensionless electric and external
902: forces acting on the particle:
903: \begin{equation}
904:   \label{eq:epsf}
905:   \vepsf := - \frac{\bF_\mathrm{el}}{2\pi\gamma r_0} , \quad
906:   \vepse := - \frac{\bF_\mathrm{ext}}{2\pi\gamma r_0} .    
907: \end{equation}
908: \\
909: 
910: \noindent
911: (iii) Finally, the following identity is a consequence of elementary
912: considerations of hydrostatics because $\Delta p$ in \eq{eq:splitPi} is
913: spatially constant:
914: \begin{equation}
915:   \label{eq:hydro}
916:   \int_{{\cal S}(\Psi)} \upd A \; \en \Delta p + \bF_{\Delta p} = 
917:   \ez A_z(\Psi) \Delta p , 
918: \end{equation}
919: where $A_z(\Psi)$ is the circular area in the $XY$ plane bounded by 
920: the circle ${\cal C}(\Psi)$.
921: 
922: Thus \eq{eq:equilibrium} can be rewritten as
923: \begin{equation}
924:   \label{eq:Spsi}
925:   \gamma \oint_{{\cal C}(\Psi)} \upd\ell \; \et \times \en = 
926:   2\pi\gamma r_0 (\vepse+\vepsf - \vepsp) 
927:   - \ez A_z(\Psi) \Delta p
928:   + \int_{\cal S_\mathrm{men}\backslash\cal S} \!\!\!\!\!\! \upd A \; 
929:   \en \Pi_\mathrm{el} . 
930: \end{equation}
931: With the notation introduced in Fig.~\ref{fig:droplet}, one has (as in
932: Sec.~\ref{sec:stress}, the subscript $_\parallel$ denotes quantities
933: evaluated at and operators acting in the undeformed spherical
934: interface, i.e., in tangent planes of the undeformed interface):
935: \begin{subequations}
936:   \label{eq:quasispher}
937: \begin{eqnarray}
938:   \en \, \upd A & = & 
939:   \left(\frac{\partial \br}{\partial \psi} \times 
940:     \frac{\partial \br}{\partial \varphi} \right) 
941:   \upd\psi \upd\varphi = 
942:   \left( 1 + \frac{u}{\rdrop} \right) 
943:   \left[\left( 1 + \frac{u}{\rdrop} \right) \er 
944:     - \nablap u \right] \upd A_\parallel , \\
945:   & & \nonumber \\
946:   \et \, \upd\ell & = & (\upd\boldsymbol{\ell}_\parallel \cdot \nabla_\parallel) \br = 
947:   \left[ 1 + \frac{u}{\rdrop} \right] \upd\boldsymbol{\ell}_\parallel +
948:   \er (\upd\boldsymbol{\ell}_\parallel \cdot \nablap u) ,
949: \end{eqnarray}
950: \end{subequations}
951: where $\upd\boldsymbol{\ell}_\parallel = \rdrop ( {\bf e}_\psi \upd
952: \psi + {\bf e}_\varphi \sin\psi \, \upd \varphi )$, $\upd A_\parallel
953: = \rdrop^2 \sin\psi \, \upd\psi \upd\varphi$,
954: and
955: \begin{equation}
956:   \nablap = \frac{{\bf e}_\psi}{\rdrop}
957:   \frac{\partial}{\partial \psi}
958:   +\frac{{\bf e}_\varphi}{\rdrop \sin\psi}
959:   \frac{\partial}{\partial\varphi}.
960: \end{equation}
961: With this notation, $A_z(\Psi) = \pi [\rdrop+u(\Psi)]^2  \sin^2\Psi$.
962: Equation~(\ref{eq:Spsi}) can be simplified under the assumption that deviations of
963: the actual droplet shape from the spherical one of radius $\rdrop$ are
964: small in the distant region $R/\rdrop \ll \Psi \leq \psi < \psi_1$, so
965: that the linearized approximation of the deformation is valid and
966: terms quadratic in the quantities $u$, $\mu$, and $\Pi_\mathrm{el}$
967: (which vanish in the unperturbed state) can be omitted \cite{ODD05a}.
968: We emphasize that this condition does not exclude large deviations
969: within the piece ${\cal S}(\Psi)$, in particular near the particle.
970: Due to rotational symmetry, the vectorial \eq{eq:Spsi} is independent
971: of the angle $\varphi$ and involves only vectors parallel to $\ez$.
972: One obtains the following ordinary differential equation for the
973: function $u(\Psi)$:
974: \begin{equation}
975:   \label{eq:1stintegral}
976:   \sin\Psi \cos\Psi \frac{\partial u}{\partial\Psi} = 
977:   r_0\, \ez\cdot(\vepse+\vepsf-\vepsp) + 
978:   \left(\frac{1}{2}\mu\rdrop - u\right) \sin^2\Psi +
979:   \frac{\rdrop^2}{\gamma} \int_{\Psi}^{\psi_1} \!\!\! \upd\psi 
980:   \; \sin\psi \cos\psi \; 
981:   \Pi_\mathrm{el}(\psi) .
982: \end{equation}
983: This expression is actually the first integral of the 
984: Young--Laplace equation (Eq.~(B7) in \rcite{ODD05a}) incorporating the
985: boundary condition at the contact line.
986: In terms of the functions $P(\psi):=\cos\psi$, $Q(\psi):=1+\cos\psi
987: \, \ln\tan(\psi/2)$ and
988: \begin{equation}
989:   S(\psi) := - \frac{\rdrop^2}{\gamma} 
990:   \int_{\psi}^{\psi_1} \!\!\! \upd s \, 
991:   \sin s \; \Pi_\mathrm{el}(s)
992:   [P(\psi) Q(s) - P(s) Q(\psi)] ,
993: \end{equation}
994: the general solution is 
995: \begin{equation}
996:   \label{eq:usolution} 
997:   u(\psi) = A P(\psi) + r_0\, \ez\cdot (\vepse+\vepsf-\vepsp) Q(\psi) +
998:   \frac{1}{2}\mu\rdrop + S(\psi),
999: \end{equation}
1000: where $\mu$ and the integration constant $A$ can be determined by the
1001: boundary condition at the plate, i.e., $\psi=\psi_1$, and the
1002: incompressibility condition of the droplet \cite{ODD05a}. In the limit
1003: $\rdrop\to\infty$ at fixed $r= \rdrop \psi$, one recovers the results
1004: of an unperturbed, planar interface \cite{DOD05}.
1005: We are particularly interested in the deformation given by
1006: \eq{eq:usolution} in the intermediate range $R/\rdrop \ll \psi \ll 1$,
1007: i.e., far from the particle and from the plate. In this range,
1008: $P(\psi) \sim 1$, $Q(\psi) \sim \ln \psi +\mbox{}$const., whereas
1009: $\Pi_\mathrm{el}(\psi)$ will decay in general as $\psi^{-n}$, implying\footnote{One finds $n=6$ in a realistic
1010: model assuming that the whole system has no net charge \cite{Wuer06a}.
1011: In the presence of a net charge,
1012: there is a monopolar electric field 
1013: far from the particle. This interesting case is beyond the scope of
1014: the present analysis, the conclusions of which only hold if
1015: $\Pi_\mathrm{el}(\psi)$ decays sufficiently fast.}
1016: $S(\psi) \sim \psi^{2-n}$ if $n>2$. Thus due to the
1017: behavior of $Q(\psi)$ there may be a logarithmically varying
1018: asymptotic deformation with the amplitude given by $r_0\, \ez\cdot
1019: (\vepse+\vepsf-\vepsp)$.
1020: 
1021: \begin{figure}
1022:   \centering{\includegraphics[width=.65\textwidth]{figure5.eps}}
1023:   \caption{The surface ${\cal S}^{(1)}$~(${\cal S}^{(2)}$) runs along
1024:     the fluid interface (full line) and the particle (dot) such that
1025:     it lies in the fluid phase exterior~(interior) to the droplet.  The
1026:     surface ${\cal S}_{plate}^{(1)}$~(${\cal S}_{plate}^{(2)}$) is that
1027:     part of the plate surface which is in contact with the fluid phase
1028:     exterior~(interior) to the droplet.  The surface ${\cal S}_L$
1029:     encloses the whole system ``particle + fluid phases'' at a
1030:     macroscopic distance $L\to\infty$ from the droplet. The surfaces
1031:     are oriented towards the exterior of the corresponding fluid volume
1032:     which they enclose.}
1033:   \label{fig:forces}
1034: \end{figure}
1035: 
1036: We now consider the situation that there is no external field acting
1037: on the system. In this case $\vepse={\bf 0}$ and the electric forces
1038: $\vepsf$ and $\vepsp$ are only due to the charge of the particle. The
1039: value of $\vepsf-\vepsp$ can be obtained by adapting the reasoning of
1040: \rcite{ODD05a} to the droplet geometry. The stress tensor in the fluid
1041: phase exterior to the droplet is given as $\mbox{}-p_1 \mathsf{1} +
1042: \stress_{\rm el}^{(1)}$, where $p_1 \mathsf{1}$ is the homogeneous,
1043: isotropic stress tensor far from the droplet and $\stress_{\rm
1044:   el}^{(1)}$ is Maxwell's stress tensor due to the electric field
1045: (modified to include the possible osmotic pressure by mobile charges,
1046: see, e.g., \rcite{SMN02}). In the same manner, $\mbox{}-(p_1 +\Delta
1047: p) \mathsf{1} + \stress_{\rm el}^{(2)}$ is the stress tensor in the
1048: interior of the droplet, where $\Delta p$ is given by \eq{eq:splitPi}.
1049: With the notations introduced in Fig.~\ref{fig:forces}, the following
1050: equations hold:
1051: \begin{subequations}
1052:   \label{eq:noforce}
1053:   \begin{equation}
1054:     \int_{
1055:       {\cal S}_{plate}^{(1)}\cup{\cal S}_L\cup{\cal S}^{(1)}
1056:     } \upd {\bf A} \cdot \stress_{\rm el}^{(1)} = {\bf 0} ,
1057:   \end{equation}
1058:   \begin{equation}
1059:     \int_{{\cal S}_{plate}^{(2)}\cup{\cal S}^{(2)}} 
1060:     \upd {\bf A} \cdot \stress_{\rm el}^{(2)} = {\bf 0} ,
1061:   \end{equation}
1062: \end{subequations}
1063: which express that the net force on the exterior fluid phase and on
1064: the interior one, respectively, vanishes in equilibrium. (The
1065: contribution to the integrals from the constant isotropic pressures
1066: $p_1$ and $p_1+\Delta p$ is zero.) On the other hand, by definition
1067: one has
1068: \begin{equation}
1069:   2\pi\gamma r_0 (\vepsp-\vepsf) = 
1070:   \mbox{}- \int_{{\cal S}^{(1)}} \upd {\bf A} \cdot 
1071:   \stress_{\rm el}^{(1)}
1072:   - \int_{{\cal S}^{(2)}} \upd {\bf A} \cdot 
1073:   \stress_{\rm el}^{(2)} .
1074: \end{equation}
1075: Combining this with Eqs.~(\ref{eq:noforce}) leads to
1076: \begin{equation}
1077:   \label{eq:nonisolation}
1078:   \vepsp-\vepsf = \frac{1}{2\pi\gamma r_0} \left[
1079:     \int_{{\cal S}_{plate}^{(1)}} \upd {\bf A} \cdot 
1080:     \stress_{\rm el}^{(1)} + 
1081:     \int_{{\cal S}_{plate}^{(2)}} \upd {\bf A} \cdot 
1082:     \stress_{\rm el}^{(2)} 
1083:   \right] ,
1084: \end{equation}
1085: where we have taken into account that the contribution of
1086: $\stress_{\rm el}^{(1)}$ over the surface ${\cal S}_L$ vanishes in the
1087: limit $L\to\infty$ because the electric field decays to zero far away
1088: from the droplet (i.e., we do not consider the possibility that there are external electric fields). 
1089: That is, $\mbox{}-(\vepsp-\vepsf)$ is actually the
1090: (dimensionless) electric force acting on the plate.
1091: The calculation of this integral requires to solve the corresponding
1092: electrostatic problem. However, on dimensional grounds one can obtain
1093: the estimate\footnote{This estimate is supported by explicit
1094:   calculations of $\Pi_\mathrm{el} = \en\cdot(\stress_{\rm
1095:     el}^{(1)}-\stress_{\rm el}^{(2)})\cdot\en$ for realistic
1096:   models \cite{Wuer06a,DaKr06a}. More precisely, the electric force
1097:   exerted on the interface is actually concentrated in a small region
1098:   of area $\sim r_0^2$ around the particle, so that one expects
1099:   $\stress_{\rm el} \sim (\gamma |\vepsp|/r_0) F(r/r_0)$, where $F$ is
1100:   a dimensionless function of order unity at distances $r \sim r_0$
1101:   from the particle and decaying $\sim (r_0/r)^n$ at distances $r \sim
1102:   \rdrop \gg r_0$. On this basis, \eq{eq:nonisolation} provides the
1103:   quoted estimate.}  $|\vepsp-\vepsf| \sim |\vepsp|
1104: (r_0/\rdrop)^{n-2}$. For an asymptotically planar interface, the
1105: logarithmically varying deformation due to nonzero values of
1106: $|\vepsp-\vepsf|$ leads to a long--ranged effective attraction (see
1107: Sec.~\ref{sec:force}), which is the reason that this mechanism has
1108: been invoked to be responsible for the apparent attraction reported in
1109: \rcite{NBHD02}.
1110: If this was the explanation, the measurements in this experiment
1111: would imply a value \cite{ODD05a} $|\vepsp-\vepsf| \sim 10^{-3}$. On
1112: the other hand, the theoretical estimate yields $|\vepsp-\vepsf| \sim
1113: 10^{-6}|\vepsp|$ with $n=6$, so that the experimental results would
1114: require $|\vepsp| \sim 10^{3}$. This large value is unlikely for
1115: realistic surface charge densities \cite{ODD05a,ODD05b}.
1116: 
1117: 
1118: This result corrects the suggestion made in \rcite{ODD05a} that
1119: $|\vepsp-\vepsf| \sim (r_0/\rdrop)^{2}$, inferred from a not
1120: applicable force balance condition. Indeed, if the deformation is
1121: small also at the contact line, the condition ``mechanical equilibrium
1122: of the particle'' can be derived from \eq{eq:1stintegral} and with an
1123: expansion in terms of $\psi_0 = r_0/\rdrop\ll 1$ leads to
1124: \begin{equation}
1125:   \left.\frac{\upd u}{\upd\psi}\right|_{\psi=\psi_0} \approx 
1126:   \rdrop \, \ez\cdot(\vepse+\vepsf) +
1127:   \psi_0 \left[
1128:     \frac{1}{2} \mu \rdrop - u(\psi_0) \right] .
1129: \end{equation}
1130: The second term, which is subdominant in the limit $\rdrop\to\infty$,
1131: is missing in Eq.~(B8) of \rcite{ODD05a}. We have cross-checked this
1132: corrected expression by deriving it also within the energy approach
1133: employed in \rcite{ODD05a}, which turns out to be algebraically much
1134: more cumbersome.
1135: 
1136: In conclusion, there persists a logarithmically varying deformation
1137: with an amplitude which is very small in the limit $\rdrop\to\infty$;
1138: this is actually a finite--size effect intrinsic to the geometry of
1139: the set-up and absent for an unbounded flat interface. However, it has the same
1140: physical origin as any logarithmically varying deformation of a flat
1141: interface,
1142: namely that the system ``particle + fluid interface'' cannot be
1143: mechanically isolated in the configuration of a droplet residing on a
1144: solid plate. In the absence of the plate one has $\vepsp-\vepsf={\bf
1145:   0}$ due to \eq{eq:nonisolation} because ${\cal S}_{plate}^{(1)}$ and
1146: ${\cal S}_{plate}^{(2)}$ are not there and the logarithmic dependence
1147: in the range $\psi \ll 1$ disappears. This conclusion corrects a
1148: recent claim of the opposite in \rcite{Wuer06b}; the relevant errors
1149: of this work are pinpointed in \rcite{DOD07a}, in particular the
1150: implementation of the boundary condition ``mechanical equilibrium of
1151: the particle''. (Reference~\onlinecite{DOD07a} represents incidentally, within the
1152: energy approach, a further confirmation of our conclusion above.)
1153: To facilitate the comparison of our calculations with \rcite{Wuer06b}
1154: we make two remarks: (i) The reasoning and results are independent of
1155: the precise functional form of the electric pressure
1156: $\Pi_\mathrm{el}(\br)$; the considerations in \rcite{Wuer06b} in this
1157: respect are thus irrelevant. (ii) As a boundary condition for fixing
1158: the droplet \rcite{Wuer06b} employs, instead of a plate at
1159: $\psi=\psi_1$ as used here, a fictional pressure field
1160: $\Pi_\mathrm{com}(\psi) \propto \cos\psi$ constraining the center of
1161: mass of the droplet \cite{MoWi93}. One can easily check that our
1162: general solution (\eq{eq:usolution}) includes this special case, as
1163: the contribution of $\Pi_\mathrm{com}(\psi)$ in \eq{eq:usolution}
1164: eliminates the singularity of $Q(\psi)$ at $\psi=\pi$ and the solution
1165: reduces to the corresponding expression in \rcite{Wuer06b}.  Thus none
1166: of these two issues affects the conclusion concerning the logarithmic
1167: dependence.
1168: 
1169: \subsection{Particle on a generally curved interface}
1170: \label{sec:curved}
1171: 
1172: If the particle is trapped at a generally curved interface ({\it
1173:   reference interface}), the electrostatic analogy can still be
1174: exploited provided there is a clear separation of length scales between the
1175: typical radius of curvature $\rdrop$ of the reference interface and
1176: the size of the region around the particle where the interfacial
1177: deformations are appreciable, say, roughly a few times the particle
1178: size $R$. Then, at distances from the particle smaller than $\rdrop$
1179: one can exploit the electrostatic analogy in order to study the small
1180: deviations of the interface from a {\it reference plane}
1181: tangent to the reference interface at some fixed point near the
1182: particle.
1183: 
1184: The deviations from the reference plane are given by the displacement field
1185: $u(\brp) = u_{\rm ref}(\brp) + \delta u(\brp)$, where $u_{\rm
1186:   ref}(\brp)$ is the deformation of the reference interface and
1187: $\delta u(\brp)$ is the additional deformation brought about by the
1188: presence of the particle. Correspondingly, the pressure field can be
1189: written as $\Pi = \Pi_{\rm ref} + \delta \Pi$.
1190: If $\ell$ is a distance from the particle beyond which the linearized
1191: theory holds (i.e., the deformation near the particle need not be
1192: small), then in the annulus $\ell<r_\parallel<\rdrop$ (with a clear
1193: separation $\ell\ll\rdrop$ so that the following dependence can be
1194: observed) the solution to the field equation can be written as
1195: \begin{eqnarray}
1196:   \label{eq:curvedu}
1197:   \delta u (\brp) & = & B_0 \ln \frac{\zeta}{r_\parallel}
1198:   + \sum_{s=1}^\infty \frac{1}{2 s} \left\{
1199:     \frac{B_s \,{\rm e}^{-i s \varphi} + B_s^* \,{\rm e}^{i s \varphi}}{r_\parallel^s}
1200:     + r_\parallel^s [ A_s \,{\rm e}^{i s \varphi} + A_s^* \,{\rm e}^{-i s \varphi}]
1201:   \right\} \nonumber \\
1202:   & & + \frac{1}{2\pi\gamma} \int_{\ell< r' <\rdrop} dA'_\parallel \;
1203:   \delta\Pi (\brp') \ln\frac{\zeta}{|\brp-\brp'|} .
1204: \end{eqnarray}
1205: The constants $A_s$ and $B_s$ are determined by the boundary conditions
1206: at $r=\ell$ and $r=\rdrop$ and account for ``virtual capillary
1207: charges'' outside the annulus. In particular, we have already seen
1208: that $2\pi\gamma B_0$ is the net bulk force on the region $r<\ell$ in
1209: the direction normal to the reference plane (and in addition to
1210: the net force in the reference state $u_{\rm ref}(\brp)$).
1211: 
1212: Two new issues arise which we have not considered so far: there is
1213: an ``external boundary'' given by the upper bound $r_\parallel=\rdrop$
1214: and there is a nonvanishing reference deformation $u_{\rm ref}(\brp)$.
1215: The first issue is not exclusive for a curved reference interface and
1216: emerges if the interfacial pressure $\Pi(\brp)$ does not decay
1217: sufficiently fast with the separation from the particle, so that in
1218: principle one cannot carry out the limit $\rdrop\to\infty$. An example has been
1219: studied in full detail in the previous subsection, where $\delta\Pi$
1220: includes a term $-\mu\gamma/\rdrop$ (see \eq{eq:splitPi}) generated by
1221: a nonlocal constraint (i.e., constant volume of the droplet). In general,
1222: the relevance of the boundary conditions at $r_\parallel=\rdrop$
1223: introduces a nonlocal ingredient preventing general
1224: statements based just on the electrostatic analogy localized around
1225: the particle.
1226: 
1227: The second issue implies an ``external electric field'',
1228: $-\nabla_\parallel u_{\rm ref}$, giving rise to a new phenomenology.
1229: To illustrate this point, we consider a reference minimal surface,
1230: $\nabla_\parallel^2 u_{\rm ref} = 0$ (so that $\Pi_{\rm ref}\equiv
1231: 0$), containing a mechanically isolated nonspherical inert particle
1232: (so that $\delta\Pi\equiv 0$) with extension $R$ much smaller than the
1233: typical curvature radius $\rdrop$ of the reference interface.
1234: As we have seen, $Q_0=Q_1=0$ (i.e., no bulk force normal to the
1235: reference plane and no torque in that plane), but $Q_2\neq 0$, so that
1236: the particle will experience a force and a torque which are given, to
1237: leading order in the small ratio $R/\rdrop$, by the coupling of this
1238: quadrupole with the ``external field'' $u_{\rm ref}$. According to
1239: \eq{eq:realcharge}, the real-valued quadrupole is characterized by the
1240: following 2nd rank tensor:
1241: \begin{equation}
1242:   \label{eq:realQ2}
1243:   \hat{Q}_2 = q_2 \left[ (\ex \ex - \ey \ey) \cos 2\theta + 
1244:     (\ex \ey + \ey \ex ) \sin 2\theta \right] ,
1245: \end{equation}
1246: where $q_2 > 0$ is the amplitude and $\theta$ is the angle which the
1247: principal axes of the quadrupole form with the coordinate axes. The
1248: origin of coordinates, $\brp={\bf 0}$, is the contact point of the
1249: reference tangent plane with the interface and must be taken at some
1250: point in or close to the particle, e.g., its center of mass. In these
1251: circumstances, the electrostatic analogy provides the force and the
1252: torque, respectively, as
1253: \begin{equation}
1254:   \bF = \frac{1}{2} \nabla_\parallel \left[ \hat{Q}_2 : 
1255:     \nabla_\parallel \nabla_\parallel u_{\rm ref} \right]_{\brp = {\bf 0}} ,
1256: \end{equation}
1257: \begin{equation}
1258:   {\bf M} = - \nabla_\parallel \times \left[ \hat{Q}_2 \cdot \nabla_\parallel u_{\rm ref} 
1259:   \right]_{\brp = {\bf 0}} ,
1260: \end{equation}
1261: after reversing the sign with respect to the electrostatic
1262: expressions, as discussed above. The most general form of the
1263: traceless Hessian matrix of the reference interface is given by
1264: \begin{equation}
1265:   \label{eq:hessian}
1266:   \nabla_\parallel\nabla_\parallel u_{\rm ref} (\br) = 
1267:   \frac{1}{\rdrop(\br)} 
1268:   \left[ (\ex \ex - \ey \ey) \cos 2\phi(\br) + 
1269:     (\ex \ey + \ey \ex ) \sin 2\phi(\br) \right] ,
1270: \end{equation}
1271: where $R_d(\br)>0$ is the absolute value of the radius of curvature
1272: and $\phi(\br)$ is the angle between the principal directions and the
1273: coordinate axes.
1274: Without loss of generality one can choose the orientation of the
1275: coordinate axes such that $\phi(\brp={\bf 0})=0$. By inserting
1276: Eqs.~(\ref{eq:realQ2}, \ref{eq:hessian}) into the previous expressions
1277: for $\bF$ and ${\bf M}$ one finally obtains
1278: \begin{equation}
1279:   \bF = q_2 \cos 2\theta \, \nabla_\parallel 
1280:   \left(\frac{1}{\rdrop}\right)_{\brp={\bf 0}} +
1281:   2 q_2 \sin 2\theta \, \left. \nabla_\parallel 
1282:     \phi \right|_{\brp={\bf 0}} ,
1283: \end{equation}
1284: \begin{equation}
1285:   {\bf M} = - \frac{2 q_2 \sin 2\theta}{\rdrop(\brp={\bf 0})}\, \ez.
1286: \end{equation}
1287: Therefore, the nonspherical particle tends to rotate so that
1288: $\theta=0,\pi$, i.e., in order to align its ``capillary quadrupole''
1289: with the principal directions of curvature of the reference, minimal interface.
1290: And when aligned like this, it is pulled in the direction of
1291: increasing curvature $1/\rdrop$.
1292: This conclusion
1293: complements the result found in \rcite{Wuer06c}, where the energy
1294: approach has been applied in order to determine the ``potential of
1295: mean force'' of a spherical\footnote{Therefore, the ``capillary
1296:   quadrupole'' is not permanent, as in the present illustrative
1297:   example, but rather induced by the ``external potential'' $u_{\rm
1298:     ref}(\brp)$ and, according to \rcite{Wuer06c}, is proportional to
1299:   the curvature $1/\rdrop$.} inert particle in a minimal
1300: surface. In a multiparticle configuration, this ``external'' force and
1301: this ``external'' torque compete with the capillary interaction
1302: between the quadrupoles (see \eq{eq:Fquad}), possibly leading to
1303: interesting phenomena concerning the 2D patterns formed by the
1304: particles.
1305: 
1306: 
1307: 
1308: \section{Summary and outlook}
1309: \label{sec:end}
1310: 
1311: We have investigated the force approach for describing colloidal particles
1312: trapped at a fluid interface.
1313: This approach has allowed us to derive a stress--tensor formulation of
1314: the interface--mediated elastic forces for an arbitrary pressure field
1315: $\Pi(\br)$ acting on the interface.  In this manner we have been able
1316: to generalize some of the results of \rcite{MDG05a} obtained only for
1317: a spatially constant pressure field. It is an interesting, open
1318: question whether this result is extendible to, e.g., membranes, for
1319: which bending rigidity as well as surface tension are relevant, and to
1320: other, more general cases considered in \rcite{MDG05a}
1321: and involving constitutive parameters beyond surface tension and
1322: bending rigidity.
1323: Based on the stress--tensor formulation we
1324: have worked out a detailed analogy between small interfacial
1325: deformations and 2D electrostatics, encompassing not only the field
1326: equation of the deformation but also the elastic forces transmitted by
1327: the interface.
1328: 
1329: We have exploited the electrostatic analogy in order to compute the dominant
1330: contribution to the interface--mediated force between two particles if
1331: they are far apart. This analogy enabled us to clarify the
1332: relationship between the energy and the force approach and to reveal the
1333: advantages and limitations of each. The definition of the effective
1334: force $F_\parallel$, which we have employed in the force approach,
1335: differs from the effective force $F_\mathrm{men}$ introduced in the
1336: energy approach via a ``potential of mean force''. However, the
1337: difference is asymptotically negligible if the interfacial
1338: deformation in the single--particle configuration is expressible as a
1339: multipole expansion (i.e., via nonvanishing ``capillary poles'').
1340: For example, if the system is not mechanically isolated or the
1341: particles are nonspherical, the force approach allows one to
1342: extend with relative ease the energy--approach result to cases in which
1343: the interfacial deformations are not small everywhere. Moreover, it justifies
1344: asymptotic pairwise additivity of the force in a multiparticle
1345: configuration.
1346: One must
1347: bear in mind that none of the two definitions $F_\parallel$ and $F_\mathrm{men}$ 
1348: is the actual force
1349: acting {\em only} on the colloid, because both take into account the
1350: force acting on the surrounding interface. This matters for discussing
1351: the dynamics of trapped particles. However, in thermal equilibrium
1352: $F_\mathrm{men}$ is the effective force according to which the
1353: equilibrium state of the particles is determined. In this situation,
1354: the energy approach, which provides $F_\mathrm{men}$, is in principle
1355: advantageous, while the force approach is more powerful (in the sense that
1356: it may facilitate or extend the range of validity of the calculations) 
1357: whenever it can be shown that $F_\parallel \approx F_\mathrm{men}$.
1358: 
1359: For the experimentally interesting system of a particle at
1360: the interface of a spherical droplet in contact with a plate
1361: (Fig.~\ref{fig:droplet}), we find that the presence of the plate
1362: breaks mechanical isolation and leads to a logarithmically varying
1363: interfacial deformation at distances $r$ from the particle in the
1364: intermediate range ``particle radius $\ll r \ll$ droplet radius'',
1365: with the amplitude of the logarithm vanishing as the droplet radius
1366: tends to infinity. Our approach has put this finite--size effect on a
1367: sound basis. Nevertheless, our numerical estimates show that 
1368: this logarithmically varying deformation is very likely too weak
1369: in order to explain the apparently long--ranged attraction observed
1370: experimentally in \rcite{NBHD02} for such a system.
1371: However, there are still open questions which we have not addressed
1372: but which are conceivably relevant for this experiment. We have
1373: assumed an electrically neutral system; but if there is a net charge,
1374: e.g., if the colloidal particle is charged but the droplet is not
1375: grounded, additional, long--ranged electric fields arise. Another
1376: interesting question is the loss of rotational symmetry which occurs
1377: if the particle is not fixed at the apex of the droplet: this might
1378: give rise to an additional force (electrostatic or capillary) pushing
1379: the particle towards the apex which, in a multiparticle configuration,
1380: could be misinterpreted as an effective attraction like the one 
1381: apparently observed also for a planar interface. 
1382: 
1383: Finally, we have discussed briefly the application of the
1384: electrostatic analogy and the associated phenomena arising when the
1385: unperturbed interface is curved. As an illustrative example, we showed
1386: that a nonspherical inert particle trapped at a minimal surface is
1387: pulled to regions in the interface with larger curvature.
1388: 
1389: 
1390: \appendix
1391: 
1392: \section{Torque balance}
1393: \label{app:dipole}
1394: 
1395: If a piece of interface ${\cal S}$ is in equilibrium, the total torque
1396: on this piece must vanish. In this case, the following condition must
1397: hold (using the same notation as in \eq{eq:equilibrium}):
1398: \begin{equation}
1399:   \label{eq:torque}
1400:   \int_{\cal S} \upd A \; (\br \times \en)\, \Pi + 
1401:   \gamma \oint_{\partial{\cal S}} \upd\ell \; \br \times 
1402:   (\et \times \en) = {\bf 0} .
1403: \end{equation}
1404: If the deviations from a flat interface are small, one can simplify
1405: \eq{eq:torque} as in Sec.~\ref{sec:stress} for the force--balance
1406: equation. To lowest order in the deformation one obtains
1407: \begin{equation}
1408:   \label{eq:torque_lin}
1409:   \left\{ 
1410:     \int_{{\cal S}_\parallel} \upd A_\parallel \; \brp \Pi + 
1411:     \gamma \oint_{\partial{\cal S}_\parallel} \upd\ell_\parallel \; 
1412:     [ \brp (\bn \cdot \nablap u) - u\bn ]
1413:   \right\} \times \ez = {\bf 0} .
1414: \end{equation}
1415: This equation implies that the expression in curly brackets vanishes
1416: because it is a vector orthogonal to $\ez$
1417: ($\brp$ and $\bn$ lie in the $XY$ plane).
1418: In the electrostatic analogy the integral over $\Pi$ corresponds to
1419: the ``capillary dipole'' ${\bf P}$ of the piece ${\cal S}$. This
1420: allows one to rewrite \eq{eq:torque_lin} as
1421: \begin{equation}
1422:   {\bf P} = \mbox{} - 
1423:   \gamma \oint_{\partial{\cal S}_\parallel} \upd\ell_\parallel \; 
1424:   [ \brp (\bn \cdot \nablap u) - u\bn ] ,
1425: \end{equation}
1426: which generalizes Gauss' theorem (\eq{eq:gauss}) by expressing the
1427: dipole of a region only in terms of the values of the deformation field and its
1428: derivatives at the boundary. On the other hand, via the general
1429: equilibrium condition in \eq{eq:torque}, the right hand side of
1430: this equality is related to the torque ${\bf M}_{\rm bulk}$ due to the
1431: bulk force:
1432: \begin{equation}
1433:   \gamma \oint_{\partial{\cal S}_\parallel} \upd\ell_\parallel \; 
1434:   [ \brp (\bn \cdot \nablap u) - u\bn ] = 
1435:   \mbox{} - \ez \times {\bf M}_{\rm bulk} .
1436: \end{equation}
1437: The validity of this expression only requires that the deformation is
1438: small at the contour $\partial {\cal S}$, where the linearization is
1439: performed, but not inside. This proofs \eq{eq:dipole}.
1440: 
1441: 
1442: \section{Multipole expansion in 2D}
1443: \label{app:multipolar}
1444: 
1445: Here we recall briefly some results concerning the multipole
1446: expansion in two dimensions. The ``potential'' $u(\br)$ created by a
1447: ``charge'' distribution $\Pi(\br)$ is given by ($\zeta$ is an
1448: arbitrary constant)
1449: \begin{equation}
1450:   \label{eq:2dfield}
1451:   u(\br) = -\frac{1}{2\pi\gamma} \int \upd A'\, \Pi(\br') 
1452:   \ln\frac{|\br-\br'|}{\zeta} =
1453:   -\frac{1}{2\pi\gamma} Re \int \upd A'\, \Pi(\br') 
1454:   \ln\frac{z-z'}{\zeta} ;
1455: \end{equation}
1456: the second equation introduces the complex variable $z=r
1457: \exp{(i\varphi)}$ in order to ease the calculations with $Re$ denoting
1458: the real part. We first consider the case that $\Pi$ has a compact
1459: support: $\Pi(\br)=0$ if $r>R$. The Taylor expansion
1460: \begin{equation}
1461:   \label{eq:logTaylor}
1462:   \ln(z-z') = \ln z - \sum_{s=1}^\infty 
1463:   \frac{1}{s} \left(\frac{z'}{z}\right)^s
1464: \end{equation}
1465: is valid in the complex domain $|z'| < |z|$. Inserting this expansion
1466: into the general expression~(\ref{eq:2dfield}) one obtains
1467: straightforwardly
1468: \begin{equation}
1469:   \label{eq:multexpansion}
1470:   u(\br) = \frac{Q_0}{2\pi\gamma} \ln \frac{\zeta}{r}
1471:   + \frac{1}{2\pi\gamma} \sum_{s=1}^\infty \frac{Q_s \,{\rm e}^{-i s \varphi} + 
1472:     Q_s^* \,{\rm e}^{i s \varphi}}{2 s r^s} ,  
1473: \end{equation}
1474: valid for $r>R$, with the complex-valued multipolar charges $Q_s$ given by
1475: \eq{eq:charge}. As can be easily deduced from this latter expression,
1476: they can be written as $Q_s = q_s \exp{(i s \theta_s)}$, where the
1477: amplitude $q_s$ is a positive real number and $\theta_s \in [0,2\pi)$
1478: is the angle by which the configuration with the charge $q_s$ is to be
1479: rotated in order to achieve a configuration with the charge $Q_s$. By
1480: using the identity $r \exp{(-i\varphi)} = \br\cdot(\ex - i \ey)$, the
1481: expansion~(\ref{eq:multexpansion}) can be rewritten in a more familiar
1482: form involving only real-valued quantities:
1483: \begin{equation}
1484:   u(\br) = \frac{\hat{Q}_0}{2\pi\gamma} \ln \frac{\zeta}{r}
1485:   + \frac{1}{2\pi\gamma} \sum_{s=1}^\infty 
1486:   \frac{\er \stackrel{(s)}{\dots} \er}{s r^s} \bullet \hat{Q}_s,  
1487: \end{equation}
1488: where $\bullet$ indicates $s$ scalar products, $\er=\br/r$, and
1489: \begin{equation}
1490:   \label{eq:realcharge}
1491:   \hat{Q}_s = Re [Q_s (\ex - i \ey) \stackrel{(s)}{\dots} 
1492:   (\ex - i \ey) ] =
1493:   q_s Re [{\rm e}^{i s \theta_s} (\ex - i \ey) \stackrel{(s)}{\dots} 
1494:   (\ex - i \ey) ] 
1495: \end{equation}
1496: are the real-valued multipolar charges.
1497: 
1498: Assume now that $\Pi\sim r^{-n}$ as $r\to\infty$, so that $Q_s$ is
1499: ill-defined for $s\geq n-2$. Nevertheless, one can still write
1500: \begin{equation}
1501:   u(\br) = \frac{Q_0}{2\pi\gamma} \ln \frac{\zeta}{r}
1502:   + \frac{1}{2\pi\gamma} \sum_{s=1}^\nu 
1503:   \frac{Q_s \,{\rm e}^{-i s \varphi} + Q_s^* \,{\rm e}^{i s \varphi}}{2 s r^s} + 
1504:   \Delta u(\br) ,  
1505: \end{equation}
1506: where $\nu$ is the largest integer such that $\nu<n-2$. This
1507: expression serves to define $\Delta u$.
1508: By using the Taylor expansion~(\ref{eq:logTaylor}) again, one can
1509: write
1510: \begin{eqnarray}
1511:   2\pi\gamma \Delta u & = & Re \int_{|z'|<|z|} \upd A'\, \Pi(\br') \,
1512:   \sum_{s=\nu+1}^\infty \frac{1}{s}\left(\frac{z'}{z}\right)^s \nonumber \\
1513:   & & + Re \int_{|z|<|z'|} \upd A'\, \Pi(\br') \,
1514:   \left[ \ln\frac{z}{z'}
1515:     + \sum_{s=1}^\infty \frac{1}{s} \left(\frac{z}{z'}\right)^s
1516:     - \sum_{s=1}^\nu \frac{1}{s} \left(\frac{z'}{z}\right)^s 
1517:   \right] .
1518: \end{eqnarray}
1519: In this form one can easily check that $\Delta u \sim r^{2-n}$ 
1520: for $r=|z|\to\infty$, and the finite multipole
1521: expansion~(\ref{eq:multipole}) holds with an extra term which is
1522: asymptotically indeed subdominant.
1523: 
1524: \begin{thebibliography}{42}
1525: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1526: \expandafter\ifx\csname bibnamefont\endcsname\relax
1527:   \def\bibnamefont#1{#1}\fi
1528: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1529:   \def\bibfnamefont#1{#1}\fi
1530: \expandafter\ifx\csname citenamefont\endcsname\relax
1531:   \def\citenamefont#1{#1}\fi
1532: \expandafter\ifx\csname url\endcsname\relax
1533:   \def\url#1{\texttt{#1}}\fi
1534: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1535: \providecommand{\bibinfo}[2]{#2}
1536: \providecommand{\eprint}[2][]{\url{#2}}
1537: 
1538: \bibitem[{\citenamefont{Ghezzi and Earnshaw}(1997)}]{GhEa97}
1539: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ghezzi}} \bibnamefont{and}
1540:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Earnshaw}},
1541:   \bibinfo{journal}{J.\ Phys.: Condensed\ Matt.} \textbf{\bibinfo{volume}{9}},
1542:   \bibinfo{pages}{L517} (\bibinfo{year}{1997}).
1543: 
1544: \bibitem[{\citenamefont{Ruiz-Garc{\'\i}a
1545:   et~al.}(1997)\citenamefont{Ruiz-Garc{\'\i}a, G\'amez-Corrales, and
1546:   Ivlev}}]{RGI97}
1547: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ruiz-Garc{\'\i}a}},
1548:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{G\'amez-Corrales}},
1549:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Ivlev}},
1550:   \bibinfo{journal}{Physica A} \textbf{\bibinfo{volume}{236}},
1551:   \bibinfo{pages}{97} (\bibinfo{year}{1997}).
1552: 
1553: \bibitem[{\citenamefont{Stamou et~al.}(2000)\citenamefont{Stamou, Duschl, and
1554:   Johannsmann}}]{SDJ00}
1555: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Stamou}},
1556:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Duschl}}, \bibnamefont{and}
1557:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Johannsmann}},
1558:   \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{5263}
1559:   (\bibinfo{year}{2000}).
1560: 
1561: \bibitem[{\citenamefont{Quesada-P\'erez
1562:   et~al.}(2001)\citenamefont{Quesada-P\'erez, Moncho-Jord\'a,
1563:   Mart{\'\i}nez-L\'opez, and Hidalgo-Alvarez}}]{QMMH01}
1564: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Quesada-P\'erez}},
1565:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moncho-Jord\'a}},
1566:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Mart{\'\i}nez-L\'opez}},
1567:   \bibnamefont{and}
1568:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hidalgo-Alvarez}},
1569:   \bibinfo{journal}{J.\ Chem.\ Phys.} \textbf{\bibinfo{volume}{115}},
1570:   \bibinfo{pages}{10897} (\bibinfo{year}{2001}).
1571: 
1572: \bibitem[{\citenamefont{Nikolaides et~al.}(2002)\citenamefont{Nikolaides,
1573:   Bausch, Hsu, Dinsmore, Brenner, Gay, and Weitz}}]{NBHD02}
1574: \bibinfo{author}{\bibfnamefont{M.~G.} \bibnamefont{Nikolaides}},
1575:   \bibinfo{author}{\bibfnamefont{A.~R.} \bibnamefont{Bausch}},
1576:   \bibinfo{author}{\bibfnamefont{M.~F.} \bibnamefont{Hsu}},
1577:   \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Dinsmore}},
1578:   \bibinfo{author}{\bibfnamefont{M.~P.} \bibnamefont{Brenner}},
1579:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gay}}, \bibnamefont{and}
1580:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Weitz}},
1581:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{420}},
1582:   \bibinfo{pages}{299} (\bibinfo{year}{2002}).
1583: 
1584: \bibitem[{\citenamefont{Tolnai et~al.}(2003)\citenamefont{Tolnai, Agod,
1585:   Kabai-Faix, Kov{\'a}cs, Ramsden, and H{\'o}rv{\"o}lgyi}}]{TAKK03}
1586: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Tolnai}},
1587:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Agod}},
1588:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kabai-Faix}},
1589:   \bibinfo{author}{\bibfnamefont{A.~L.} \bibnamefont{Kov{\'a}cs}},
1590:   \bibinfo{author}{\bibfnamefont{J.~J.} \bibnamefont{Ramsden}},
1591:   \bibnamefont{and}
1592:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{H{\'o}rv{\"o}lgyi}},
1593:   \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{107}},
1594:   \bibinfo{pages}{11109} (\bibinfo{year}{2003}).
1595: 
1596: \bibitem[{\citenamefont{Fern\'andez-Toledano
1597:   et~al.}(2004)\citenamefont{Fern\'andez-Toledano, Moncho-Jord\'a,
1598:   Mart{\'\i}nez-L\'opez, and Hidalgo-Alvarez}}]{FMMH04}
1599: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Fern\'andez-Toledano}},
1600:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moncho-Jord\'a}},
1601:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Mart{\'\i}nez-L\'opez}},
1602:   \bibnamefont{and}
1603:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hidalgo-Alvarez}},
1604:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{20}},
1605:   \bibinfo{pages}{6977 } (\bibinfo{year}{2004}).
1606: 
1607: \bibitem[{\citenamefont{Chen et~al.}(2005)\citenamefont{Chen, Tan, Ng, Ford,
1608:   and Tong}}]{CTNF05}
1609: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Chen}},
1610:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Tan}},
1611:   \bibinfo{author}{\bibfnamefont{T.-K.} \bibnamefont{Ng}},
1612:   \bibinfo{author}{\bibfnamefont{W.~T.} \bibnamefont{Ford}}, \bibnamefont{and}
1613:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tong}},
1614:   \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{95}},
1615:   \bibinfo{pages}{218301} (\bibinfo{year}{2005}).
1616: 
1617: \bibitem[{\citenamefont{Nicolson}(1949)}]{Nico49}
1618: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Nicolson}},
1619:   \bibinfo{journal}{Proc.\ Cambridge Philos.\ Soc.}
1620:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{288} (\bibinfo{year}{1949}).
1621: 
1622: \bibitem[{\citenamefont{Chan et~al.}(1981)\citenamefont{Chan, {Henry Jr.}, and
1623:   White}}]{CHW81}
1624: \bibinfo{author}{\bibfnamefont{D.~Y.~C.} \bibnamefont{Chan}},
1625:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{{Henry Jr.}}},
1626:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{White}},
1627:   \bibinfo{journal}{J.\ Coll.\ Interface Sci.} \textbf{\bibinfo{volume}{79}},
1628:   \bibinfo{pages}{410} (\bibinfo{year}{1981}).
1629: 
1630: \bibitem[{\citenamefont{Megens and Aizenberg}(2003)}]{MeAi03}
1631: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Megens}} \bibnamefont{and}
1632:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Aizenberg}},
1633:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{424}},
1634:   \bibinfo{pages}{1014} (\bibinfo{year}{2003}).
1635: 
1636: \bibitem[{\citenamefont{Foret and W\"urger}(2004)}]{FoWu04}
1637: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Foret}} \bibnamefont{and}
1638:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W\"urger}},
1639:   \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{92}},
1640:   \bibinfo{pages}{058302} (\bibinfo{year}{2004}).
1641: 
1642: \bibitem[{\citenamefont{Oettel et~al.}(2005{\natexlab{a}})\citenamefont{Oettel,
1643:   Dom{\'\i}nguez, and Dietrich}}]{ODD05a}
1644: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1645:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1646:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1647:   \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{71}},
1648:   \bibinfo{pages}{051401} (\bibinfo{year}{2005}{\natexlab{a}}).
1649: 
1650: \bibitem[{\citenamefont{Oettel et~al.}(2005{\natexlab{b}})\citenamefont{Oettel,
1651:   Dom{\'\i}nguez, and Dietrich}}]{ODD05b}
1652: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1653:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1654:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1655:   \bibinfo{journal}{J.\ Phys.: Condensed\ Matt.} \textbf{\bibinfo{volume}{17}},
1656:   \bibinfo{pages}{L337} (\bibinfo{year}{2005}{\natexlab{b}}).
1657: 
1658: \bibitem[{\citenamefont{W{\"u}rger and Foret}(2005)}]{WuFo05}
1659: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}} \bibnamefont{and}
1660:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Foret}}, \bibinfo{journal}{J.
1661:   Phys. Chem. B} \textbf{\bibinfo{volume}{109}}, \bibinfo{pages}{16435}
1662:   (\bibinfo{year}{2005}).
1663: 
1664: \bibitem[{\citenamefont{Dom{\'\i}nguez
1665:   et~al.}(2005)\citenamefont{Dom{\'\i}nguez, Oettel, and Dietrich}}]{DOD05}
1666: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1667:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
1668:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1669:   \bibinfo{journal}{J.\ Phys.: Condensed Matt.} \textbf{\bibinfo{volume}{17}},
1670:   \bibinfo{pages}{S3387} (\bibinfo{year}{2005}).
1671: 
1672: \bibitem[{\citenamefont{Oettel et~al.}(2006)\citenamefont{Oettel,
1673:   Dom{\'\i}nguez, and Dietrich}}]{ODD06}
1674: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1675:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1676:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1677:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{22}},
1678:   \bibinfo{pages}{846} (\bibinfo{year}{2006}).
1679: 
1680: \bibitem[{\citenamefont{Danov et~al.}(2006)\citenamefont{Danov, Kralchevsky,
1681:   and Boneva}}]{DKB06}
1682: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}},
1683:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
1684:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~P.}
1685:   \bibnamefont{Boneva}}, \bibinfo{journal}{Langmuir}
1686:   \textbf{\bibinfo{volume}{22}}, \bibinfo{pages}{2653} (\bibinfo{year}{2006}).
1687: 
1688: \bibitem[{\citenamefont{Dom{\'\i}nguez
1689:   et~al.}(2007{\natexlab{a}})\citenamefont{Dom{\'\i}nguez, Oettel, and
1690:   Dietrich}}]{DOD06a}
1691: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1692:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
1693:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1694:   \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{127}},
1695:   \bibinfo{pages}{204706} (\bibinfo{year}{2007}{\natexlab{a}}).
1696: 
1697: \bibitem[{\citenamefont{Segel}(1987)}]{Sege77}
1698: \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Segel}},
1699:   \emph{\bibinfo{title}{Mathematics applied to continuum mechanics}}
1700:   (\bibinfo{publisher}{Dover, New York}, \bibinfo{year}{1987}).
1701: 
1702: \bibitem[{\citenamefont{Kralchevsky et~al.}(1993)\citenamefont{Kralchevsky,
1703:   Paunov, Denkov, Ivanov, and Nagayama}}]{KPDI93}
1704: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
1705:   \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Paunov}},
1706:   \bibinfo{author}{\bibfnamefont{N.~D.} \bibnamefont{Denkov}},
1707:   \bibinfo{author}{\bibfnamefont{I.~B.} \bibnamefont{Ivanov}},
1708:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Nagayama}},
1709:   \bibinfo{journal}{J. Coll. Interface Sci.} \textbf{\bibinfo{volume}{155}},
1710:   \bibinfo{pages}{420} (\bibinfo{year}{1993}).
1711: 
1712: \bibitem[{\citenamefont{M{\"u}ller et~al.}(2005)\citenamefont{M{\"u}ller,
1713:   Deserno, and Guven}}]{MDG05a}
1714: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{M{\"u}ller}},
1715:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Deserno}}, \bibnamefont{and}
1716:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Guven}},
1717:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{482}
1718:   (\bibinfo{year}{2005}).
1719: 
1720: \bibitem[{\citenamefont{Aveyard et~al.}(2000)\citenamefont{Aveyard, Clint,
1721:   Nees, and Paunov}}]{ACNP00}
1722: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Aveyard}},
1723:   \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Clint}},
1724:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Nees}}, \bibnamefont{and}
1725:   \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Paunov}},
1726:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{16}},
1727:   \bibinfo{pages}{1969} (\bibinfo{year}{2000}).
1728: 
1729: \bibitem[{\citenamefont{Loudet et~al.}(2006)\citenamefont{Loudet, Yodh, and
1730:   Pouligny}}]{LYP06}
1731: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Loudet}},
1732:   \bibinfo{author}{\bibfnamefont{A.~G.} \bibnamefont{Yodh}}, \bibnamefont{and}
1733:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Pouligny}},
1734:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{97}},
1735:   \bibinfo{pages}{018304} (\bibinfo{year}{2006}).
1736: 
1737: \bibitem[{\citenamefont{Smalyukh et~al.}(2004)\citenamefont{Smalyukh,
1738:   Chernyshuk, Lev, Nych, Ognysta, Nazarenko, and Lavrentovich}}]{SCLN04}
1739: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Smalyukh}},
1740:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Chernyshuk}},
1741:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Lev}},
1742:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Nych}},
1743:   \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Ognysta}},
1744:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Nazarenko}},
1745:   \bibnamefont{and}
1746:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Lavrentovich}},
1747:   \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{93}},
1748:   \bibinfo{pages}{117801} (\bibinfo{year}{2004}).
1749: 
1750: \bibitem[{\citenamefont{Kralchevsky et~al.}(2001)\citenamefont{Kralchevsky,
1751:   Denkov, and Danov}}]{KDD01}
1752: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
1753:   \bibinfo{author}{\bibfnamefont{N.~D.} \bibnamefont{Denkov}},
1754:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}},
1755:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{17}},
1756:   \bibinfo{pages}{7694} (\bibinfo{year}{2001}).
1757: 
1758: \bibitem[{\citenamefont{Paunov}(1998)}]{Paun98}
1759: \bibinfo{author}{\bibfnamefont{V.~N.} \bibnamefont{Paunov}},
1760:   \bibinfo{journal}{Langmuir} \textbf{\bibinfo{volume}{14}},
1761:   \bibinfo{pages}{5088} (\bibinfo{year}{1998}).
1762: 
1763: \bibitem[{\citenamefont{Kleman and Lavrentovich}(2003)}]{KlLa03}
1764: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kleman}} \bibnamefont{and}
1765:   \bibinfo{author}{\bibfnamefont{O.~D.} \bibnamefont{Lavrentovich}},
1766:   \emph{\bibinfo{title}{Soft Matter Physics}} (\bibinfo{publisher}{Springer,
1767:   New York}, \bibinfo{year}{2003}).
1768: 
1769: \bibitem[{\citenamefont{Hurd}(1985)}]{Hurd85}
1770: \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Hurd}},
1771:   \bibinfo{journal}{J.\ Phys.\ A: Math.\ Gen.} \textbf{\bibinfo{volume}{18}},
1772:   \bibinfo{pages}{L1055} (\bibinfo{year}{1985}).
1773: 
1774: \bibitem[{\citenamefont{Danov and Kralchevsky}(2006)}]{DaKr06a}
1775: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}} \bibnamefont{and}
1776:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
1777:   \bibinfo{journal}{J. Coll. Interface Sci.} \textbf{\bibinfo{volume}{298}},
1778:   \bibinfo{pages}{213} (\bibinfo{year}{2006}).
1779: 
1780: \bibitem[{\citenamefont{Dom{\'i}nguez et~al.}()\citenamefont{Dom{\'i}nguez,
1781:   Frydel, and Oettel}}]{DFO08}
1782: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'i}nguez}},
1783:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Frydel}}, \bibnamefont{and}
1784:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1785:   \bibinfo{journal}{\pre}
1786:   \textbf{\bibinfo{volume}{77}}, \bibinfo{pages}{020401(R)} (\bibinfo{year}{2008}).
1787: 
1788: \bibitem[{\citenamefont{Oettel et~al.}(2008)\citenamefont{Oettel,
1789:   Dom{\'\i}nguez, Tasinkevych, and Dietrich}}]{ODTD07}
1790: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1791:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1792:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tasinkevych}},
1793:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1794:   \bibinfo{journal}{unpublished}  (\bibinfo{year}{2008}).
1795: 
1796: \bibitem[{\citenamefont{Danov et~al.}(2004)\citenamefont{Danov, Kralchevsky,
1797:   and Boneva}}]{DKB04}
1798: \bibinfo{author}{\bibfnamefont{K.~D.} \bibnamefont{Danov}},
1799:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Kralchevsky}},
1800:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~P.}
1801:   \bibnamefont{Boneva}}, \bibinfo{journal}{Langmuir}
1802:   \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{6139} (\bibinfo{year}{2004}).
1803: 
1804: \bibitem[{\citenamefont{Lehle et~al.}(2006)\citenamefont{Lehle, Oettel, and
1805:   Dietrich}}]{LOD06}
1806: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Lehle}},
1807:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
1808:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1809:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{174}
1810:   (\bibinfo{year}{2006}).
1811: 
1812: \bibitem[{\citenamefont{Lehle and Oettel}(2007)}]{LeOe07}
1813: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Lehle}} \bibnamefont{and}
1814:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}},
1815:   \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{011602} (\bibinfo{year}{2007}).
1816: 
1817: \bibitem[{\citenamefont{Fournier and Galatola}(2002)}]{FoGa02}
1818: \bibinfo{author}{\bibfnamefont{J.-B.} \bibnamefont{Fournier}} \bibnamefont{and}
1819:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Galatola}},
1820:   \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{65}},
1821:   \bibinfo{pages}{031601} (\bibinfo{year}{2002}).
1822: 
1823: \bibitem[{\citenamefont{van Nierop et~al.}(2005)\citenamefont{van Nierop,
1824:   Stijnman, and Hilgenfeldt}}]{VSH05}
1825: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{van Nierop}},
1826:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Stijnman}},
1827:   \bibnamefont{and}
1828:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Hilgenfeldt}},
1829:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{72}}, \bibinfo{pages}{671}
1830:   (\bibinfo{year}{2005}).
1831: 
1832: \bibitem[{\citenamefont{Shestakov et~al.}(2002)\citenamefont{Shestakov,
1833:   Milovich, and Noy}}]{SMN02}
1834: \bibinfo{author}{\bibfnamefont{A.~I.} \bibnamefont{Shestakov}},
1835:   \bibinfo{author}{\bibfnamefont{J.~L.} \bibnamefont{Milovich}},
1836:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Noy}},
1837:   \bibinfo{journal}{J. Coll. Interface Sci.} \textbf{\bibinfo{volume}{247}},
1838:   \bibinfo{pages}{62} (\bibinfo{year}{2002}).
1839: 
1840: \bibitem[{\citenamefont{W{\"u}rger}(2006{\natexlab{a}})}]{Wuer06a}
1841: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}},
1842:   \bibinfo{journal}{Eur. J. Phys. E} \textbf{\bibinfo{volume}{19}},
1843:   \bibinfo{pages}{5} (\bibinfo{year}{2006}{\natexlab{a}}).
1844: 
1845: \bibitem[{\citenamefont{W{\"u}rger}(2006{\natexlab{b}})}]{Wuer06b}
1846: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}},
1847:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{978}
1848:   (\bibinfo{year}{2006}{\natexlab{b}}).
1849: 
1850: \bibitem[{\citenamefont{Dom{\'\i}nguez
1851:   et~al.}(2007{\natexlab{b}})\citenamefont{Dom{\'\i}nguez, Oettel, and
1852:   Dietrich}}]{DOD07a}
1853: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dom{\'\i}nguez}},
1854:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oettel}}, \bibnamefont{and}
1855:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dietrich}},
1856:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{77}}, \bibinfo{pages}{68002}
1857:   (\bibinfo{year}{2007}{\natexlab{b}}).
1858: 
1859: \bibitem[{\citenamefont{Morse and Witten}(1993)}]{MoWi93}
1860: \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Morse}} \bibnamefont{and}
1861:   \bibinfo{author}{\bibfnamefont{T.~A.} \bibnamefont{Witten}},
1862:   \bibinfo{journal}{\epl} \textbf{\bibinfo{volume}{22}}, \bibinfo{pages}{549}
1863:   (\bibinfo{year}{1993}).
1864: 
1865: \bibitem[{\citenamefont{W{\"u}rger}(2006{\natexlab{c}})}]{Wuer06c}
1866: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{W{\"u}rger}},
1867:   \bibinfo{journal}{\pre} \textbf{\bibinfo{volume}{74}},
1868:   \bibinfo{pages}{041402} (\bibinfo{year}{2006}{\natexlab{c}}).
1869: 
1870: 
1871: 
1872: \end{thebibliography}
1873: 
1874: 
1875: 
1876: 
1877: \end{document}
1878: 
1879: 
1880: