0807.1852/ms.tex
1: % author- ID 51475 ms-ID E08071
2: %%%%%%%%%%%%%%%%%%%%%%% file template.tex %%%%%%%%%%%%%%%%%%%%%%%%%
3: %
4: % This is a template file for The European Physical Journal
5: %
6: % Copy it to a new file with a new name and use it as the basis
7: % for your article
8: %
9: %%%%%%%%%%%%%%%%%%%%%%%% Springer-Verlag %%%%%%%%%%%%%%%%%%%%%%%%%%
10: %
11: \begin{filecontents}{leer.eps}
12: %!PS-Adobe-2.0 EPSF-2.0
13: %%CreationDate: Mon Jul 13 16:51:17 1992
14: %%DocumentFonts: (atend)
15: %%Pages: 0 1
16: %%BoundingBox: 72 31 601 342
17: %%EndComments
18: 
19: gsave
20: 72 31 moveto
21: 72 342 lineto
22: 601 342 lineto
23: 601 31 lineto
24: 72 31 lineto
25: showpage
26: grestore
27: %%Trailer
28: %%DocumentFonts: Helvetica
29: \end{filecontents}
30: %
31: %\documentclass[epj,referee]{svjour}
32: \documentclass[epj]{svjour}
33: 
34: \newcommand{\bea}{\begin{eqnarray}}
35: \newcommand{\eea}{\end{eqnarray}}
36: \newcommand{\spaceint}[2]{\int_{#1} d^3 #2 \;}
37: \newcommand{\vect}[1]{\mathbf{#1}}
38: \newcommand{\vat}{V^{\rm att}}
39: \newcommand{\di}{\displaystyle}
40: 
41: \usepackage{amsmath,amssymb}
42: \usepackage{latexsym}
43: \usepackage{epsfig,color}
44: \usepackage[normalem]{ulem}
45: 
46: %\setlength{\parindent}{0cm}
47: %\setlength{\parskip}{0.5cm}
48: %\setlength{\textwidth}{17cm}
49: %\setlength{\textheight}{24cm}
50: %\addtolength{\oddsidemargin}{-2.5cm}
51: %\addtolength{\evensidemargin}{-2.5cm}
52: %\addtolength{\topmargin}{-2.5cm}
53: 
54: \begin{document}
55: 
56: %\newlength{\mylen}
57: %\setlength{\mylen}{\textwidth}
58: %\addtolength{\mylen}{-1cm}
59: 
60: 
61: %\preprint{{\LARGE \bf DRAFT}}
62: 
63: \title{Effective interactions of colloids on nematic films}
64: 
65: \author{M. Oettel\inst{1}, A. Dom\'\i nguez\inst{2}, M. Tasinkevych\inst{3}
66:  \and S. Dietrich\inst{3}}
67: 
68: \institute{
69:  Johannes--Gutenberg--Universit\"at Mainz, Institut f{\"ur} Physik,
70:   WA 331, D--55099 Mainz, Germany 
71: \and
72:  F\'\i sica Te\'orica, Universidad de Sevilla, Apdo.1065, E-41080
73:   Sevilla, Spain
74: \and 
75:  Max-Planck-Institut f\"ur Metallforschung, Heisenbergstr. 3,
76:   D-70569 Stuttgart and
77:  Institut f\"ur Theoretische und Angewandte Physik, Universit\"at Stuttgart,
78:              Pfaffenwaldring 57, D-70569 Stuttgart, Germany
79: }
80: 
81: \date{Received: / Revised version:}
82: % The correct dates will be entered by Springer
83: %
84: 
85: 
86: \abstract{
87:   The elastic and capillary interactions  between a pair of colloidal particles
88:   trapped on top of a nematic film are studied theoretically 
89:   for large separations $d$. The elastic interaction is repulsive and 
90:   of quadrupolar type, varying as $d^{-5}$.
91:   For macroscopically thick films, the capillary interaction is likewise
92:   repulsive and proportional to $d^{-5}$ as a consequence
93:   of mechanical isolation of the system comprised of the colloids and the interface.
94:   A finite film thickness introduces a nonvanishing force on the system
95:   (exerted by the substrate supporting the film)
96:   leading to logarithmically varying capillary attractions. 
97:   However, their strength turns out to be too small to 
98:   be of importance for the recently observed pattern formation of 
99:   colloidal droplets on
100:   nematic films.
101: }
102: 
103: \PACS{ {82.70.Dd}{Colloids} \and
104:       {68.03.Cd} {Surface tension and related phenomena} \and
105:    {61.30.-v}{Liquid Crystals}
106:      } % end of PACS codes
107: 
108: \maketitle
109: 
110: \section{Introduction}
111: 
112: The interactions of colloidal particles trapped at fluid interfaces have been found to differ
113: significantly from the corresponding interactions in bulk solvents. This has been studied
114: mostly for electrically charged particles trapped at interfaces with water. On 
115: one hand, the presence of the interface gives rise to direct dipolar electrostatic
116: repulsions between the colloids (see Refs.~\cite{Pie80,Ave02,Par07} for some  experimental evidence),
117: on the other hand deformations of the interface may induce longer--ranged
118: capillary attractions (briefly reviewed in Refs.~\cite{Zen06,Bre7r,Oet08}) which 
119: is possibly the source of 
120: pattern formation observed
121: in various experiments \cite{Ghe97,Gar98a,Que01,Gom05,Che06}. 
122: (See, however, Ref.~\cite{Fer04} for an alternative
123: explanation due to interface impurities.)
124: 
125: Recently \cite{Lav04,Lav07}, the experimental observation of ordered structures of glycerol 
126: droplets bound to a 
127: nematic--air interface has been reported and attributed to
128: an effective pair potential between the colloids which contains a repulsive,
129: elastic part due to director deformations in the supporting
130: nematic film and an attractive, capillary part which is long--ranged and mediated 
131: by logarithmically 
132: varying deformations of the nematic--isotropic interface caused by the droplets.   
133: The schematic setup of this experiment is depicted in
134: Fig.~\ref{fig:nem_exp}. According to Ref.~\cite{Lav04}, the colloidal particles 
135: experience an 
136: upward force caused by elastic forces due to director deformations in the supporting
137: nematic film. 
138: %and the resulting surface mediated colloid--colloid interaction can be evaluated 
139: %in the superposition approximation. 
140: This upward force on the colloids is supposed to give rise to
141: the aforementioned logarithmically varying  interface deformation.
142: Applying a superposition approximation for the
143: deformation field, one can show that the ensuing effective capillary interaction
144: potential between two colloids is likewise varying logarithmically. 
145: This is similar to the flotation interaction
146: of mm--sized particles at fluid interfaces for which the force on the colloids is caused
147: by gravity (see, e.g., Ref.~\cite{Kra00}) and also parallels the tentative 
148: explanation given for the experimentally observed attractions
149: between sub-$\mu$m charged colloids at a water-oil interface \cite{Nik02}
150: (for the controversy around this explanation see 
151: Refs.~\cite{Meg03,Nik03,For04,Kra04,Oet05,Oet05a,Kra05,Dan06}). 
152: 
153: \begin{figure}[t]
154:  \begin{center}
155:   \epsfig{file=nem_exp.eps, width=\columnwidth}
156:  \end{center}
157:  \caption{ \label{fig:nem_exp}
158:  Schematic setup of the experiment reported in Ref.~\cite{Lav04}. 
159:  Colloidal glycerol drops ($R = 1 \dots 7$ $\mu$m) are trapped at the surface of a   
160:  thick nematic film ($h \approx 60$ $\mu$m). The director field in the
161:  nematic film is sketched by the black lines and their dotted interpolations. 
162:  Nematic anchoring at the glycerol substrate at the bottom of the film
163:  as well as on the surface of the glycerol drops 
164:  is parallel, whereas it is perpendicular (homeotropic) 
165:  at the deformed nematic--air interface. Each
166:  drop is necessarily accompanied by a topological defect ($\otimes$). 
167:  }
168: \end{figure}
169: 
170: However, it is now well established \cite{Meg03,Nik03,For04,Oet05,Oet05a,Kra05,Dom06}
171: %some of us have shown recently \cite{Oet05} 
172: that interface deformations and effective colloidal interactions varying logarithmically only arise in experimental systems which are not isolated
173: mechanically. For mechanically isolated systems it can be shown \cite{Oet05b,Wue05,Dom05,Dom07}
174: that both the interface deformation around  a single  colloid and the effective
175: interaction between two of them  are shorter--ranged and the latter cannot be 
176: calculated reliably within the
177: superposition approximation. 
178: 
179: In the following we will extend the arguments presented in 
180: Refs.~\cite{Oet05,Oet05b,Wue05,Dom05,Dom07}
181: to systems with colloids at nematic interfaces. We will show that mechanical isolation
182: of the system ``nematic film -- colloid -- air" can be violated through a subtle interplay
183: between the finite thickness of the film and the anchoring conditions at the colloids
184: and at the nematic interfaces with the substrate and with the  air, respectively. 
185: However, for 
186: experimental conditions as the ones described in 
187: Ref.~\cite{Lav04} a quantitative estimate of the strength of the ensuing 
188: logarithmic attraction between the colloids yields that these attractions
189: are unobservably small. Therefore it seems likely that this kind of asymptotic capillary forces
190: cannot be invoked as a relevant mechanism to account for the observations 
191: reported in Ref.~\cite{Lav04}.  
192: 
193: The manuscript is organized as follows:
194: In Sec.~\ref{sec:coarse} the coarse--grained model for the nematic
195: phase is introduced which will serve as the basis for all subsequent
196: calculations. In Subsec.~\ref{sec:thick} we study the case of an
197: infinitely thick nematic film. First, we compute the asymptotic form of
198: the director field and the ensuing elastic force between two
199: particles. Then we calculate the effective force arising from the
200: deformation of the fluid--nematic interface caused by the elastic
201: stresses. In Subsec.~\ref{sec:thin} we consider a nematic film of finite
202: thickness, which models more closely the experimental setup described
203: in Ref.~\cite{Lav04}; for such a system we extend the above calculations 
204: to the two opposite cases
205: of perpendicular and parallel anchoring of the director field at the
206: substrate surface.  In Sec.~\ref{sec:conclusion} we
207: discuss our results.
208: 
209: %that for a  
210: %mechanically isolated experimental system (which is reasonable to assume since
211: %gravity is unimportant for $\mu$m--sized particles) an acting force on the colloid
212: %is necessarily accompanied by a stress on the interface in such a way that 
213: %for small forces the
214: %resulting interface deformation is {\em not} logarithmic but short--ranged:
215: %if the stress field on the interface around one colloid $\Pi(r)$ decays like
216: %$r^{-n}$ the interface deformation $u(r)$ decays like $r^{-n+2}$.   
217: %Consequently, the superposition approximation predicts a short--ranged, 
218: %{\em repulsive} surface mediated potential $V_{\rm men}(d)$ for colloids at distance
219: %$d$ which varies as $d^{-n}$. Although we have investigated 
220: %only charged colloids in more detail, the arguments are quite general and can be applied
221: %to a nematic interface as well which we will demonstrate below. 
222: 
223: 
224: \section{Coarse--grained model}
225: \label{sec:coarse}
226: 
227: In view of the mesoscopic length scales involved we describe the bulk part of the 
228: nematic free energy associated with the director deformations
229: in terms of the Frank free energy expression within the one--coupling approximation
230: \cite{deG74}
231: \bea
232:   {\cal F}^{\rm b}_{\rm ne} &=&  \spaceint{V_{\rm ne}}{r} f^{\rm b} (\vect r) \; \nonumber \\
233:  \label{eq:f_lc}
234:            &=& \frac{K}{2}\spaceint{V_{\rm ne}}{r} \left[ (\nabla \cdot \vect n)^2
235:    + (\nabla \times \vect n)^2 \right]   \\
236:     &=& \frac{K}{2}\spaceint{V_{\rm ne}}{r} \nabla n_{i}\cdot \nabla n_{i} + \nonumber \\
237:  & &   
238:   \frac{K}{2}\spaceint{V_{\rm ne}}{r} \nabla \cdot[ \vect n (\nabla \cdot \vect n) -(\vect n \cdot \nabla) \vect n] \;. \nonumber
239: \eea
240: $V_{\rm ne}$ denotes the volume occupied by the nematic film, $\vect n$ is the
241: director field ($\vect n^2 =1$), and the constant $K$ is of the order of $10^{-11}$ N
242: \cite{deG74}. 
243: The  total divergence term in the last line of Eq.~(\ref{eq:f_lc}) exhibits 
244: the so--called ``$K_{24}$--structure" and is unimportant for the bulk equations describing
245: the equilibrium configuration.
246: The surface free energies associated with the interfaces with air and substrate, 
247: respectively, (see Fig.~\ref{fig:nem_exp}) 
248: are described in terms of the Poulini expression \cite{Lav03}: 
249: \bea
250:  \label{eq:fs_lc}   
251:    {\cal F}^{\rm s}_{\rm ne} &=& \frac{W_1}{2} \int_{A_{\rm air-ne}}
252:   \!\!\!\!\!\!\!\!\!\!\!\! dA \;
253:    (\vect n \cdot \vect e_A)^2 + 
254: %  \nonumber \\ & &
255:    \frac{W_2}{2} 
256:    \int_{A_{\rm sub-ne} \bigcup A_{\rm coll-ne}}
257:   \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! dA \;
258:    (\vect n \cdot \vect e_A)^2  \;.
259: \eea 
260: %{\tt New things included from here till Eq.(3)}
261: Here, $\vect e_A$ denotes the local surface normal unit vector pointing outwards from the
262: film or the colloid. Normal alignment is favored
263: for $W_i<0$ and parallel alignment for $W_i>0$. 
264: %
265: %and the director adopts the favored alignment over a length of the
266: %order of $K/|W_i|$.
267: %
268: Typically one has
269: $|W_i| \sim 10^{-5}$ N/m \cite{Lav03} so that the length scale $K/|W_i| \sim 1$ $\mu$m is smaller
270: than the range of droplet radii investigated in Ref.~\cite{Lav04}.
271: Thus, in the ``strong anchoring'' limit which we shall consider, %this case, 
272: the effect of the boundary terms is so strong that
273: as a first approximation it amounts to  
274: fixing the angle between the director and the surface normal. 
275: We shall adopt $W_1<0$ (normal alignment at the nematic--air interface) and $W_2>0$ 
276: (parallel alignment at the nematic--glycerol interfaces). Some consequences of 
277: deviations from the strong anchoring limit will be discussed in App.~\ref{sec:app1}.
278: 
279: These surface contributions to the free energy (``wetting energies'') 
280: are small corrections to the 
281: surface tensions which are mainly due to dispersion interactions.
282: We denote these non--nematic contributions to the surface tensions as  
283: $\gamma_1$ (colloid--air surface
284: tension), $\gamma_2$ (substrate--nematic surface tension) and $\gamma'$ (nematic--air
285: surface tension). Typically, these surface tensions are of the order of  $10^{-2}$ N/m.
286: Therefore, they are much larger than the constants $W_i$ which determine the nematic
287: contributions to the surface tension. We see that due to 
288: the above anchoring conditions the surface tension of the substrate--nematic interface
289: carries no nematic contributions due to $\vect n \cdot \vect e_A=0$ (Eq.~(\ref{eq:fs_lc})) 
290: whereas in the strong anchoring limit the full nematic--air surface tension 
291: is $\gamma= \gamma' + W_1/2 \approx \gamma'$ due to $\vect n \cdot \vect e_A=1$ 
292: (Eq.~(\ref{eq:fs_lc})).\footnote{
293: A genuine contribution to the surface tension of a nematic interface
294: arises if one takes into account the variation of the nematic tensorial order
295: parameter through the interface, described by, e.g., the Landau--de Gennes
296: free energy functional (generalizing Eq.~(\ref{eq:f_lc})). 
297: The magnitude of these contributions
298: can be estimated by the surface tension of the interface between the nematic and
299: the isotropic phase of a liquid crystal. Typically, such a surface tension
300: is also $O(10^{-5}$ N/m) $\sim |W_i|$ and therefore small compared to the dispersion
301: force contribution. Furthermore, a distorted director structure in the {\em bulk}
302: may also give rise to surface energy contributions on the boundaries. For an 
303: example see Ref.~\cite{deG74}, p. 131 and p. 174. Also in this case it can be
304: argued that the corresponding contributions to the surface tension do not exceed
305: $|W_i|$.  } 
306: %
307: %Ref.~\cite{Lav04} identifies the force associated to the anchoring
308: %``wetting energy'' $\propto W_2$ with the ultimate origin of the
309: %capillary attraction. We notice, however, that this argument is
310: %incomplete: if the particle is only partially wetted by the nematic
311: %phase (contact angle $\theta \in (0,\pi)$, as found experimentally),
312: %this force will be counteracted by the force associated to the
313: %``wetting energy'' of the particle--air interface. Therefore, the
314: %particle can be trapped at an undeformed fluid interface
315: %%without the need of a deformation of the meniscus 
316: %and remain in mechanical equilibrium \cite{Pie80} (this effect is
317: %described by the second term in the free energy functional in
318: %Eq.~(\ref{eq:F}) below). In the present case, however, there are also elastic
319: %forces pulling at the interface. Our goal is the computation of the
320: %interfacial deformation and the ensuing effective capillary
321: %interaction due to these elastic forces.
322: 
323:   
324: The canonical stress tensor $\pi_{ij}$ associated with the free energy 
325: expression in Eq.~(\ref{eq:f_lc})
326: is given by 
327: \bea
328:  \label{eq:stresstensor}
329:  \pi_{ij} = \frac{\partial{ f^{\rm b}}}{\partial n_{k,j}}\;n_{k,i} - \delta_{ij}\;f^{\rm b}
330: \eea
331: where $n_{k,i} =\partial n_k/\partial x_i$ (summation over $k$). 
332: The total stress tensor $\Pi_{ij}$ is obtained by
333: adding the contribution of the isotropic pressure $p$:
334: \bea
335:  \Pi_{ij} = \pi_{ij} - \delta_{ij}\;p\;.
336: \eea
337: 
338: \subsection{Macroscopically thick nematic film}
339: \label{sec:thick}
340: 
341: First we consider the limiting case $h \to \infty$ (i.e., very thick nematic films (see 
342: Fig.~\ref{fig:nem_exp})).\footnote{This was 
343: implicitly assumed also  by the authors of  Ref.~\cite{Lav04} in discussing Fig.~2 
344: therein.} Due to 
345: the small values of the elastic coupling constant, $K \ll \gamma\, R$, 
346: %(see Fig.~\ref{fig:nem_exp})
347: and
348: of the anchoring energy, $|W_i| \ll \gamma$, the equilibrium 
349: configuration of  a single colloid at the nematic--air interface deviates 
350: only slightly from
351: the reference configuration depicted in Fig.~\ref{fig:ref}. In this latter configuration, 
352: the interface
353: is flat and the colloid is positioned such that the contact angle fulfills Young's law
354: $\cos \theta = (\gamma_1-\gamma_2)/\gamma$.
355: 
356: \begin{figure}[t]
357:  \begin{center}
358:   \epsfig{file=ref.eps, width=\columnwidth}
359:  \end{center}
360:   \caption{\label{fig:ref}
361:      In the reference configuration the whole system is divided into
362:       volumes $V_1$ and $V_2$. Volume $V_2$ (enclosed by the
363:       upper dashed curve) includes the air and the glycerol drop and volume $V_1$ (enclosed by
364:       the lower dashed curve) includes the nematic.
365:       The arrows indicate the
366:       direction in which the surfaces (including the infinitesimally
367:       displaced ones) are oriented: $S$
368:       encloses the whole system,  $S_{\rm men}$ is the interface
369:       between the nematic phase and air (acting as a meniscus), and $S_1$ is the 
370:       interface between the colloidal
371:       drop and the nematic phase. The director field and the topological 
372:       defect ($\otimes$) are
373:       indicated as in Fig.~\ref{fig:nem_exp}. The radius of the three--phase
374:       contact line is denoted by $\rho_0=R\sin\theta$ where $\theta$ is the contact angle
375:       of the air--nematic interface with the colloid of radius $R$. $\rho$ denotes
376:       the lateral distance from the vertical symmetry axis of the colloidal drop.}
377: \end{figure}
378: 
379: The total force on the whole system reads (the
380: superscript $^{+(-)}$ denotes evaluation on the positive (negative)
381: side of the oriented surface, i.e., on the side the arrows in Fig.~\ref{fig:ref}
382: [which indicate the surface normals] point to (do not point to)):
383: \bea
384:  \oint_S  d{\bf A} \cdot {\mathbf \Pi} &=& \int_{V_1\bigcup V_2} 
385:   \!\!\!\!\!\! dV \; 
386:    (\nabla \cdot {\mathbf \Pi}) + \int_{S_{\rm men}\bigcup S_1} 
387:   \!\!\!\!\!\! d{\bf A} \cdot 
388:   ({\mathbf \Pi}^+ - {\mathbf \Pi}^-) \nonumber \\
389:   &=& - \int_{S_{\rm men}}  dA \; (\pi_{zz}+p_{\rm air}-p)  \, {\bf e}_z  + 
390:   \nonumber \\ &&
391:   \int_{S_1} d{\bf A}\cdot \left[{\boldsymbol \pi}+(p_{\rm air}-p){\bf 1}\right]
392:    \nonumber \\
393:   &=& - \int_{S_{\rm men}}  dA \; \pi_{zz}  \, {\bf e}_z  + 
394:   \int_{S_1} d{\bf A}\cdot {\boldsymbol \pi}  \; .
395:  \label{eq:Ftot}
396: \eea
397: In obtaining this equation we have applied Gauss' theorem. Furthermore
398: we have used the relation
399: $\nabla \cdot {\mathbf \Pi}=0$ in volumina $V_1$ and $V_2$ which is valid because 
400: the reference configuration is taken 
401: to be in force equilibrium. This also implies that the  isotropic pressures above the
402: interface ($p_{\rm air}$) and below it ($p$) are equal 
403: and that the director configuration is given by the 
404: corresponding Euler--Lagrange equilibrium equations following from the functional
405: in Eq.~(\ref{eq:f_lc}).
406: Since at the interface $S_1$ the colloidal drop is rigidly attached to the liquid crystal,
407: we can identify the vertical force $F$ on the colloid and the total force $F_\pi$
408: on the air--nematic interface by
409: \bea
410:  \label{eq:Fdef}
411:   F\;{\bf e}_z &=& \int_{S_1} d{\bf A}\cdot {\boldsymbol \pi}
412: \eea
413: and 
414: \bea 
415:  \label{eq:Fpi}
416:   F_\pi &=& -\int_{S_{\rm men}}  dA \; \pi_{zz} \; ,
417: \eea
418: respectively.
419: Mechanical isolation of the system means that the total force  
420: $\oint_S  d{\bf A} \cdot {\mathbf \Pi}$ acting  on it 
421: is zero which leads to
422: \bea
423:  \label{eq:f_eq_pi}
424:   F&=& F_\pi\;.
425: \eea 
426: 
427: For a given force on the colloid and a given stress on the interface, the
428: interface deformation relative to the reference configuration can be determined perturbatively.
429: To that end, we summarize briefly those results of Ref.~\cite{Oet05} which are pertinent
430: also for the present system. 
431: With the introduction of the  two small, dimensionless parameters 
432: \bea
433:   \label{eq:epsdef}
434:   \varepsilon_F = - F/(2\pi\,\rho_0\,\gamma) \quad \mbox{and} \quad
435:   \varepsilon_\pi = - F_\pi/(2\pi\,\rho_0\,\gamma)
436: \eea
437: one can  expand (up to second order in
438: $\varepsilon_F$ and $\varepsilon_\pi$) the free energy difference ${\cal F}$  
439: associated with the interface deformation $u(\rho\ge \rho_0)$ 
440: around a single colloid (see Fig.~\ref{fig:ref}) 
441: and with a vertical shift $\Delta h$ which is the difference of the  colloid center position
442: relative to that in  the reference configuration: 
443: \bea
444:   {\cal F} & \simeq & 
445:   2 \pi \gamma \int_{\rho_0}^{\infty} d\rho \; \rho \left[ \frac{1}{2}\left(\frac{d u}{d \rho} \right)^2 +
446:     \frac{u^2}{2 \lambda^2} + \frac{1}{\gamma} \pi_{zz} \, u \right] +
447:  \nonumber \\ &&
448:   \label{eq:F}
449:   \pi \gamma [ u(\rho_0) - \Delta h ]^2 -
450:   F \Delta h \;.
451: \eea
452: Here, $\lambda=(\gamma/(g\bar\rho_{\rm m}))^{1/2}$ is the capillary length associated 
453: with the interface where $g$ is the gravitational constant and $\bar\rho_{\rm m}$ is the
454: mass density of the nematic phase. This expression
455: for the free energy contains all surface free energy changes relative to the reference
456: configuration involving the interfaces between air, nematic,
457: or colloid. It also  contains the contributions due to  volume forces acting on the nematic 
458: (associated with $\lambda$)
459: and the energy change of the colloid upon vertical shifts (for further details see
460: Ref.~\cite{Oet05}).
461: Note that 
462: to leading (quadratic) order in $\varepsilon_\pi,\varepsilon_F$
463: the free energy change of the nematic due to the shifted interface
464: and due to a change in the director configuration with respect
465: to the reference configuration is captured 
466: by the term $\propto \int \pi_{zz}\,u$.
467: (The analogous textbook argument for electrostatics \cite{Sch98} can be 
468: easily generalized
469: to the nematic case described by the free energy expression in Eq.~(\ref{eq:f_lc}).) 
470: 
471: Minimizing ${\cal F}$ with respect to $u(\rho)$ and $\Delta h$ and 
472: focussing on  the regime $\rho \ll \lambda$ yields 
473: \bea
474:   \label{eq:sol1}
475:   u (\rho) \simeq \rho_0 (\varepsilon_\pi - \varepsilon_F) \ln\frac{C \lambda}{\rho}
476:   + \frac{1}{\gamma} \int_{\rho}^{+\infty} \!\!\!\! d\sigma \; \sigma \, \pi_{zz}(\sigma) \ln \frac{\sigma}{\rho} \;, \quad
477: \eea
478: with $C \simeq 1.12$. We see that in the case of an isolated system
479: $(\varepsilon_\pi= \varepsilon_F)$ the logarithmic part of $u(\rho)$ vanishes. The 
480: second term on the rhs of Eq.~(\ref{eq:sol1}) leads to 
481: $u(\rho \to \infty) \propto \rho^{-n+2}$
482: if $\pi_{zz} \propto \rho^{-n}$ and thus describes a shorter--ranged power--law
483: decay of the interface deformation. 
484: 
485: The absence of logarithmic deformations for an isolated system has been derived here
486: under certain simplifying conditions (small interfacial deformation everywhere, rotational symmetry).
487: %which allows using the reference configuration. 
488: In App.~\ref{sec:app2} we demonstrate that this conclusion
489: holds in general.
490: % less restrictive case that the interface deformation is small
491: % only far away from the particles.
492:  
493: 
494: \subsubsection{Asymptotic director configuration and elastic force between colloids}
495: 
496: 
497: %Next we derive
498: %the exponent $n$ of the asymptotic stress field $\pi_{zz}$ for our simple 
499: %director model, defined by Eq.~(\ref{eq:f_lc}) and 
500: %perpendicular anchoring at the interface as realized experimentally in 
501: %Ref.~\cite{Lav04}. 
502: In Refs.~\cite{Pou97,Lub98} it has been shown that a colloidal drop 
503: immersed in the bulk of a liquid crystal
504: is accompanied by a single counterdefect such that the total topological charge is zero
505: (here, the volume occupied by the colloid contains a topological charge which
506: may be represented by a virtual defect inside the colloid) and
507: the asymptotic behavior of the director field is of dipolar character.
508: Based on similar considerations we shall show that for a colloidal drop located at the
509: air--nematic interface the boundary conditions for that interface impose a 
510: quadrupole--like  asymptotic 
511: behavior of the director field. Macroscopically far from the colloid the director
512: is oriented parallel to the $z$ axis.
513: Accordingly, at large but finite distances $r$ the director is
514: given by $\vect n(\vect r) \simeq (n_1,n_2,1-O(n_1^2,n_2^2))$ and the bulk
515: free energy corresponding to Eq.~(\ref{eq:f_lc}) is  given by
516: \bea
517:  \label{eq:fasy}
518:   {\cal F}^{\rm b}_{\rm ne} \simeq \frac{K}{2} \spaceint{V_{\rm ne}}{r} 
519:     \left( \sum_{i=1,2} \left( \nabla n_i 
520:     \right)^2 + O (n_i^4) \right)\;.
521: \eea
522: Here we have discarded the total divergence term in the free energy expression
523: (\ref{eq:f_lc}). It is unimportant for the bulk equations and it adds a mere
524: constant to the free energy because the director is anchored normally at 
525: the boundary (the nematic--air interface).
526: Thus for each component
527: $i=1,2$ the equilibrium director field fulfills the Laplace equation
528: \bea
529:  \label{eq:lapl}
530:    \Delta n_i = 0\;.
531: \eea
532: Analogously to electrostatics, the asymptotic solution for $n_i$ can be expanded
533: in terms of multipoles. To this end we  consider the reference configuration in
534:  Fig.~\ref{fig:ref}. 
535: We choose as the origin of the coordinate system the center 
536: of the circle formed by the planar three--phase contact line. 
537: The solution for the director field has to fulfill the following requirements:
538: ($i$) rotational covariance around the $z$-axis\footnote{If $\bf D$ specifies 
539: the transformation matrix for such a rotation then this
540: requirement is given by $\vect n (\bf D\cdot \vect r) = \bf D \cdot \vect n
541: (\vect r)$.} and ($ii$) $n_i(x,y,z=0)=0$.
542: Analyzing the multipole {\em ansatz} (with $\vect r=(r_1, r_2, r_3)$)
543: %\alpha,\;\alpha=1,2,3)$)
544: \bea
545:  \label{eq:es_mul}
546:   n_i &=& q_i\,\frac{1}{r} + \sum_{\alpha=1}^3 P_{i\alpha}\,\frac{r_\alpha}{r^3} +  
547:       \sum_{\alpha,\beta=1}^3 Q_{i\alpha\beta}\,\frac{r_\alpha\,r_\beta}{r^5}
548:  + \dots
549: \eea
550: it follows that rotational covariance requires $q_i=0$, 
551: $ P_{i\alpha} = P\,\delta_{i\alpha} + P_{\rm mag}\,\epsilon_{i\alpha 3}$ 
552: and $Q_{i\alpha\beta}=Q'_\beta\,\delta_{i\alpha} + 
553: Q'_{{\rm mag},\beta}\,\epsilon_{i\alpha 3}$ ($\epsilon_{ijk}$ is the Levi-Civit\`a tensor).
554: $P_{\rm mag}$ and $Q'_{{\rm mag},\beta}$ are dipole and quadrupole moments,
555: respectively, for a director field of ``magnetic" type, i.e., for
556: which ${\rm div}\,n_i=0$ holds. 
557: The boundary condition ($ii$) at the interface with the air further imposes 
558: $P=P_{\rm mag}=0$ and
559: $Q'_\beta= Q\, \delta_{\beta3}$, $Q'_{{\rm mag},\beta}= Q_{\rm mag}\, 
560: \delta_{\beta3}$. It appears to be difficult geometrically to match
561: the asymptotic solution of ``magnetic" type  with a solution near the
562: colloid which obeys parallel anchoring at the colloid surface. 
563: %\textcolor{blue}{\sout{Furthermore,
564: %if we assumed weak anchoring ($|W|R/K \ll 1$), the magnetic
565: %quadrupole would vanish identically as can be shown in a perturbative
566: %calculation along the lines of Ref.~\cite{Ruh97}. }}
567: Therefore we 
568: discard the magnetic quadrupole,
569:  i.e., the leading asymptotic term is given by
570: the remaining, ``electric" quadrupole term:
571: \bea
572:    \label{eq:quadrupole}
573:   n_i  &=& Q\; \frac{z\,r_i}{r^5} +\dots \qquad (z \equiv r_3) \;. 
574: \eea 
575: This is at variance with Ref.~\cite{Lav04} 
576: (see Eq.~(3) therein and the considerations in the paragraph above that equation 
577: which assume
578: a dipole field) but it is consistent with the analysis in Ref.~\cite{Lub98}. 
579: Dimensional analysis yields $Q=O(R^3)$ \cite{Lub98}. 
580: Note that we have derived the asymptotic behavior of the director field at the
581: interface using strong anchoring at the interface ($n_1=n_2=0$). 
582: In App.~\ref{sec:app1}
583: we discuss corrections to strong anchoring which are, however, of subleading
584: character and leave the leading behavior (Eq.~(\ref{eq:quadrupole})) unchanged.
585: 
586: The asymptotic elastic interaction between two colloids in the bulk at distance $d$ accompanied
587: by a quadrupolar director deformation has been analyzed in Ref.~\cite{Ruh97}
588: (for weak anchoring) and in Refs.~\cite{Ram96,Lub98} (using a coarse--graining
589: method, applicable also for strong anchoring) yielding identical results. For the present configuration (distance
590: vector perpendicular to the asymptotic director) the elastic
591: potential is repulsive  and varies as
592: \begin{equation}
593:   V_{\rm el} \propto \frac{K\,Q^2}{d^5} \propto \gamma \rho_0^2\,
594:    \varepsilon_F\left(\frac{\rho_0}{d}\right)^5\;. 
595: \end{equation}
596: We have used that the dimensionless force parameter $\varepsilon_F$ is proportional to
597: $Q^2$
598: which actually follows from Eqs.~(\ref{eq:Fdef}), (\ref{eq:f_eq_pi}), (\ref{eq:epsdef}), and  
599: (\ref{eq:pizz_single}) below.
600: Note that we have simply extrapolated the {\em bulk} results for two interacting
601: colloids which cause asymptotically quadrupolar deformations of the director.
602:  This appears
603: to be reasonable because the asymptotic director field in the nematic phase for 
604: the interface
605: problem is  precisely that of the bulk solution in the lower half plane
606: and because the bulk solution is antisymmetric with respect to $z \to -z$, thus
607: respecting the boundary condition $n_i(x,y,z=0)=0$. However, the precise
608: numerical value of the quadrupole moment $Q$ might be rather different for the case
609: of the colloid trapped at the interface as compared to the bulk case.
610: 
611: \subsubsection{Asymptotic behavior of the stress on the interface and meniscus--induced 
612: effective potential between colloids}
613: 
614: The asymptotic behavior of the stress tensor component $\pi_{zz}$ at the interface
615: follows from inserting 
616: Eq.~(\ref{eq:quadrupole}) into Eq.~(\ref{eq:stresstensor}):
617: \bea
618:  \left. \pi_{zz}\right|_{\rm interface} &=&
619: %  \left.\frac{K}{2} \sum_{i=1}^2 \left(n_{i,z}n_{i,z}-n_{i,r_1}n_{i,r_1}-n_{i,r_2}n_{i,r_2}
620:   \left.\frac{K}{2} \sum_{i=1}^2 \left(n_{i,z}^2-n_{i,r_1}^2-n_{i,r_2}^2
621:    \right)\right|_{z=0} 
622:  \nonumber \\  
623:  \label{eq:pizz_single}
624:   &\stackrel{r\to\infty}{\longrightarrow} &  \frac{K}{2}\,Q^2 \,\frac{1}{\rho^8} \;, 
625: \eea
626: ($\rho^2=r_1^2+r_2^2$).
627: Consequently the interface
628: deformation around a single colloid for a mechanically isolated system obeys
629: $u(\rho\to\infty)\propto \rho^{-6}$ (see Eq.~(\ref{eq:sol1})). 
630: 
631: For the problem of two identical colloids  located at ${\boldsymbol \rho}_1$ and 
632: ${\boldsymbol \rho}_2$ (vectors are defined in the interface plane $z=0$) a distance
633: $d=|{\boldsymbol \rho}_1-{\boldsymbol \rho}_2| \gg R$ apart the expression for the
634: free energy is a straightforward generalization of the one for the single--colloid
635: free energy given in Eq.~(\ref{eq:F}):
636: \bea
637:   \label{eq:F2}
638:   {\hat {\cal F}} &= &
639:   \gamma \int_{S_{\rm men}} \!\!\!\!\!\! d^2 \rho \;
640:   \left[  \frac{|\nabla \hat{u}|^2}{2} + \frac{\hat{u}^2}{2 \lambda^2} -
641:     \frac{\hat{\pi}_{zz}}{\gamma} \, \hat{u} \right] +
642:  \\ \nonumber & &
643:    \sum_{i=1,2} \left\{
644:     \frac{\gamma}{2 \rho_{0}} \oint_{\partial S_i} \!\!\! d\ell \; [\Delta \hat{h}_i - \hat{u}]^2
645:     - \hat{F}_i \Delta \hat{h}_i \right\} .
646: % \nonumber
647: \eea
648: Here, $\hat{F}_i$ denotes the force on colloid $i$ and  $\Delta \hat{h}_i$ is the relative
649: position of its center. The integration domain $S_{\rm men}$ is the whole interface plane
650: except for the two circular disks bordered by the (reference configuration) contact lines  
651: $\partial S_i$. The meniscus--induced effective potential 
652: is the difference between the equilibrium free energy of the two colloids at distance $d$
653: and their free energy at macroscopic distance:
654: \bea
655:   V_{\rm men}(d) = {\hat {\cal F}}_{\rm eq} (d) - {\hat {\cal F}}_{\rm eq} (d\to\infty) \;.
656: \eea
657: As before, minimization with respect to
658: $\hat u(\boldsymbol \rho)$ and $\Delta \hat h_i$ renders the equilibrium free energy.
659: 
660: The behavior of $V_{\rm men}(d)$ has been analyzed  in detail in 
661: Refs.~\cite{Oet05,Dom07}. Here we summarize these results as far as they are relevant
662: for the present problem.
663: The interfacial stress $\hat \pi_{zz}$ % and the interface deformation field $\hat u$ 
664: may be decomposed generally as
665: \bea
666:   \hat \pi_{zz}(\boldsymbol \rho)& = &\pi_{zz}(|\boldsymbol \rho-\boldsymbol \rho_1|) + \pi_{zz}(|\boldsymbol \rho-\boldsymbol \rho_2|) 
667:   + 2\,\pi_{zz, {\rm m}}(\boldsymbol \rho)
668:  \nonumber \\
669:   \label{eq:pihat}
670:  & \equiv & \pi_{zz,1}+\pi_{zz,2}+2\,\pi_{zz, {\rm m}} \;.% \\
671: %  \label{eq:uhat}
672: %  \hat u(\boldsymbol \rho)& = &u(|\boldsymbol \rho-\boldsymbol \rho_1|) + u(|\boldsymbol \rho-\boldsymbol \rho_2|) + u_{\rm m}(\boldsymbol \rho)
673: %  \equiv u_1+u_2+u_{\rm m} \;. 
674: \eea
675: Here, $\pi_{zz, i}$ denotes the stress %and  interface deformation field
676: around  colloid $i$ which pertains to the problem of a single colloid. 
677: To quadratic order the asymptotic director field around 
678: two colloids is given by the superposition of the components $n_i$ of the
679: single--colloid solutions and thus to this order we recover the decomposition
680: in Eq.~(\ref{eq:pihat}) with the mixed
681: component of stress field $\pi_{zz, {\rm m}}$ given by
682: \bea
683:   \pi_{zz, {\rm m}} =
684:  \frac{K}{2}\;Q^2\; \frac{(\boldsymbol \rho-\boldsymbol \rho_1) \cdot (\boldsymbol \rho-\boldsymbol \rho_2)}
685: {|\boldsymbol \rho-\boldsymbol \rho_1|^5\; |\boldsymbol \rho-\boldsymbol \rho_2|^5} \;.
686: \eea
687: 
688: It turns out that  
689: for a system under an external force $(\varepsilon_\pi \not = \varepsilon_F)$
690: the mixed term $\pi_{zz, {\rm m}}$ does not contribute to the leading
691: term in $V_{\rm men}$. %and $u_{\rm m}$. 
692: Thus in the case of a non--vanishing external force this leading
693: contribution to $V_{\rm men}$ is obtained by a superposition 
694: {\em ansatz} which  consists in 
695: approximating the interfacial deformation and the total %deformation and 
696: stress field by the sum of the respective single--colloid quantities only
697: ($\hat \pi_{zz}\approx \pi_{zz,1}+\pi_{zz,2}$,
698:  $\hat u \approx u_1+u_2$) \cite{Oet05,Dom07}:
699: \bea
700:  \label{eq:vmen_noniso}
701:  V_{\rm men}(\rho_0 \ll d \ll \lambda) &\simeq&   - 2\pi\;\gamma\; \rho_0^2\,
702:    (\varepsilon_\pi - \varepsilon_F)^2 \, \ln \frac{C\lambda}{d}  \qquad
703:   \\ && \nonumber
704:      (\varepsilon_\pi \not = \varepsilon_F)\;.
705: \eea
706: For an isolated system  $(\varepsilon_\pi= \varepsilon_F)$ and for the stress
707: given in Eq.~(\ref{eq:pizz_single}) the superposition approximation 
708: $\hat \pi_{zz}\approx \pi_{zz,1}+\pi_{zz,2}$
709: yields $V_{\rm men} \propto \varepsilon_F^2/d^8$ as the dominant term. 
710: However, this is not the leading term, which rather stems from 
711: $\pi_{zz,{\rm m}}$. This term has two peaks around the colloid centers and 
712: therefore close to the colloids it can be approximated by
713: \bea
714:   \label{eq:pimmpeak}
715:   \pi_{zz,{\rm m}} \approx \frac{K \,Q^2}{2\,d^4} \sum_{i=1}^2 (-1)^i
716:   \frac{ \vect e_d \cdot(\boldsymbol \rho-\boldsymbol \rho_i)}
717:   {|\boldsymbol \rho-\boldsymbol \rho_i|^5} 
718: \eea
719: where $\vect e_d = (\boldsymbol \rho_2-\boldsymbol \rho_1)/d$.
720: %This mixed stress gives rise to mixed contributions $u_{\rm m}$ which are also
721: %centered around the colloids and whose overall amplitude is determined by the 
722: %prefactor of $\pi_{zz,{\rm m}}$ and thus we have $u_{\rm m} \propto d^{-4}$.
723: As discussed in Ref.~\cite{Dom07}, the qualitative behavior of  $V_{\rm men}(d)$  
724: is captured by the integral over $\pi_{zz,{\rm m}}$:
725: \bea
726:  \label{eq:vmen_iso}
727:   V_{\rm men}(d) &\propto & \gamma\rho_0^3\,\varepsilon_F \int_{S_{\rm men}} \!\!\! d^2\rho \, 
728:    \pi_{zz,{\rm m}}(\boldsymbol \rho) 
729:   \propto  \gamma\rho_0^2\,\varepsilon_F^2\,\left(\frac{\rho_0}{d}\right)^5 \qquad
730:  \\ && \nonumber
731:   (\varepsilon_\pi = \varepsilon_F)\;.
732: \eea
733: We note that this form of the mixed stress and thus the leading behavior of 
734: $V_{\rm men}(d)$
735: is formally analogous to the contributions of the
736: electric field components parallel to the interface in the case of charged colloids
737: \cite{Dom07}.  
738: The meniscus--induced potential $V_{\rm men}$ is repulsive and 
739: falls off asymptotically\footnote{Superficially one would expect a 
740: leading decay 
741: $V_{\rm men}\propto d^{-4}$ as displayed in Eq.~(\ref{eq:pimmpeak}). However, due to
742: the geometric factor in the numerator of Eq.~(\ref{eq:pimmpeak}), this apparent
743: leading order vanishes upon integration.} $\propto
744: d^{-5}$, as does the likewise repulsive elastic 
745: potential $V_{\rm el}$. We note that 
746: $V_{\rm el} \propto |\varepsilon_F|$ and $V_{\rm men} \propto \varepsilon_F^2$, so that
747: the meniscus--induced potential is small compared to the elastic one for
748: the parameters of the experiment reported in Ref.~\cite{Lav04}.
749: 
750: 
751: 
752: In order to explain the experimentally observed attractions, in Ref.~\cite{Lav04} 
753: a perturbative 
754: picture similar to the one presented
755: above was suggested, but the effect of the interface stress $\pi_{zz}$ was
756: neglected completely. (A similar error has been made in Ref.~\cite{Nik02} which
757: the authors of Ref.~\cite{Lav04} refer to.) 
758: %{\tt The following sentence has been rewritten:} 
759: In a heuristic way, only an ``upward" force on the colloid (perpendicular to the interface)
760:  associated with
761: the anchoring ``wetting" energy at the nematic--particle interface
762: has been invoked in Ref.~\cite{Lav04}, neglecting
763: the force on the interface described by $\pi_{zz}$.
764: % which, as we have seen, is incomplete.
765: % a change of the director
766: % field between the reference and the deformed configuration which can be seen
767: % to be of higher order
768: % in the parameters $\varepsilon_\pi, \varepsilon_F$. 
769: In this way, the unbalance of the force
770: gives rise to a logarithmic term in the meniscus
771: deformation and in the meniscus--induced potential. However, mechanical isolation 
772: (i.e., force balance) renders
773: the meniscus--induced potential actually repulsive and shorter--ranged 
774: (see Eq.~(\ref{eq:vmen_iso})), apparently in contrast to the experimental results.
775: A logarithmically varying potential can only arise if mechanical isolation is violated (see App.~\ref{sec:app2}). 
776: Below we shall investigate whether a net force on the system ``colloid and interface"
777: may appear if the
778: thickness of the nematic phase is finite, as it is the case in the actual experiment. 
779: 
780:  
781: 
782: 
783: 
784: 
785: \subsection{Finite thickness of the nematic film}
786: \label{sec:thin}
787: 
788: In our discussion of a finite film thickness of the nematic phase we shall consider two cases:
789: \begin{enumerate}
790: \item  The anchoring of the nematic director at the surface of the bottom substrate is 
791: perpendicular as it is the case at the upper interface with the air. 
792: This case bears a strong formal resemblance to  charged
793: colloids on water surfaces which have been  discussed in Refs.~\cite{Oet05,Dom07}. 
794: \item The anchoring at the bottom substrate surface is parallel (as in the experiment reported in
795: Ref.~\cite{Lav04}). At large lateral distances from the colloids, this leads to a director
796: field which gradually rotates from parallel orientation at the bottom substrate to the
797: perpendicular orientation at the upper interface. 
798: \end{enumerate}
799: %{\tt new intermediate summary}
800: %In combination with the normalization
801: %condition ${\bf n}^2=1$ this leads to a strong net force
802: %and consequently to a rather strong interface deformation. This effect is peculiar 
803: %to this experimental system.  
804: 
805: For both cases the total force on the system -- comprising air, nematic film, 
806: colloid, and the substrate -- must be zero. This leads to
807: (see Fig.~\ref{fig:ref1}(a) and compare with Eq.~(\ref{eq:Ftot}))
808: \bea
809:  0 &=& \oint_S  d{\bf A} \cdot {\mathbf \Pi} \nonumber \\
810:    &=& \int_{V_1\bigcup V_2\bigcup V_2'} \!\!\!\!\!\!\!\!\! dV \; 
811:    (\nabla \cdot {\mathbf \Pi}) + \int_{S_{\rm men}\bigcup S_1\bigcup S_{\rm sub}} \!\!\!\!\!\!\!\!\! d{\bf A} \cdot 
812:   ({\mathbf \Pi}^+ - {\mathbf \Pi}^-) \nonumber \\
813:   &=& - \int_{S_{\rm men}}  dA \; (\pi_{zz}+p_{\rm air}-p)  \, {\bf e}_z  +
814:  \label{eq:Ftot1} \\
815:  && \int_{S_1} \!\!\!d{\bf A}\cdot \left[{\boldsymbol \pi}+(p_{\rm air}-p){\bf 1}\right]
816:   + \int_{S_{\rm sub}} \!\!\!\!\!\!dA \; (\pi_{zz}+p_{\rm sub}-p)  \, {\bf e}_z  \; .
817:  \nonumber
818: \eea
819: For reasons which will become clear in the  discussion of  the second anchoring case, 
820: we consider the isotropic pressures in the substrate and air, $p_{\rm sub}$ and
821: $p_{\rm air}$, respectively, not necessarily to be equal to the pressure $p$ in the 
822: nematic film. The second equation in Eq.~(\ref{eq:Ftot1}) follows from the equilibrium
823: condition $\nabla \cdot {\mathbf \Pi}=0$ which holds  in all volumina.
824: 
825: \subsubsection{Perpendicular anchoring at both interfaces}
826: 
827: \label{sec:aligned}
828: 
829: \begin{figure}[t]
830:  \begin{center}
831:   \epsfig{file=ref1.eps, width=\columnwidth}
832:  \end{center}
833:   \caption{\label{fig:ref1}
834:     The reference configuration for a nematic film (a) differs from the reference 
835:     configuration
836:     shown in Fig.~\ref{fig:ref} by the addition of the substrate volume $V_2'$.
837:     The interface between the  substrate and the nematic film is denoted by 
838:     $S_{\rm sub}$.
839:     Panel (b) shows the string of image quadrupoles $ Q_b=Q$ which are needed 
840:     to fulfill the boundary conditions for perpendicular anchoring of the director 
841:     field (originating
842:     from the colloid quadrupole $Q$) at both interfaces confining the nematic film. 
843:     The distance between any two nearest neighbor quadrupoles is $2h$.
844:    }
845: \end{figure}
846: 
847: As discussed above, the presence of the  colloid asymptotically generates a 
848: quadrupolar director field  
849: which fulfills the boundary condition at the nematic--air interface.
850: In order to fulfill the
851: boundary condition at the substrate--nematic interface, an image quadrupole
852: of the same strength $Q$ is needed which, however, leads to a violation of the
853: nematic--air interface boundary condition and requires a second image quadrupole etc.
854: Continuation of this process leads to a string of image quadrupoles as 
855: depicted in Fig.~\ref{fig:ref1}(b). This string of quadrupoles
856: generates a stress field ${\boldsymbol \pi}$ which vanishes for large lateral distances.
857: Therefore the isotropic pressures must be equal in all volumina:
858: $p_{\rm sub}=p_{\rm air}=p$. From Eq.~(\ref{eq:Ftot1}) one finds that the difference
859: between the force $F$ on the colloid and the integrated stress 
860: $F_\pi$ at the nematic--air interface,
861: \bea
862:  \label{eq:df1}
863:   F- F_\pi &=& \Delta F = - \int_{S_{\rm sub}}  dA \; \pi_{zz}\;,
864: \eea
865: is given by the integrated stress over the substrate surface, i.e., the total force
866: on all quadrupoles above the substrate surface exerted by the image quadrupoles
867: in the substrate. Expressed in terms of the force $F_{Q-Q}$ between two quadrupoles
868: at distance $2h$, we find
869: \bea
870:  -\Delta F / F_{Q-Q} & =&   \sum_{n=1}^\infty \frac{1}{n^6} + \sum_{n=2}^\infty 
871:  \frac{1}{n^6} + \dots \nonumber \\ 
872:     &  = &  \sum_{n=1}^\infty \frac{1}{n^5} = \zeta(5) \approx 1.04  \;.
873: \eea
874: In this equation, the first sum is the total force (divided by $F_{Q-Q}$) on the first
875: quadrupole above the substrate exerted by all quadrupoles in the substrate,
876: the second sum is the total force (divided by $F_{Q-Q}$) on the second quadrupole
877: above the substrate etc. 
878: For the force between two quadrupoles at distance $2h$ we find, 
879: using Eqs.~(\ref{eq:quadrupole}),
880: (\ref{eq:stresstensor}), and (\ref{eq:df1}):
881: \bea
882:  \label{eq:fqq}
883:    F_{Q-Q} & =& \frac{5}{6}\pi\,K\,\frac{Q^2}{h^6}\;.
884: \eea
885: Thus mechanical isolation for the system ``colloid and \linebreak nematic--air interface" 
886: is violated and the total force $\Delta F$ on this system 
887: is (up to the factor 1.04) given by the quadrupole force $F_{Q-Q}$.
888: Nevertheless the magnitude of the corresponding induced logarithmic capillary potential
889: (see Eq.~(\ref{eq:vmen_noniso}) with $\varepsilon_\pi - \varepsilon_F = \Delta F / (2\pi\gamma\rho_0)$) is small, because $Q \sim R^3$ due to dimensional arguments
890: \cite{Lub98}.
891: For parameters similar
892: to the ones appearing in Ref.~\cite{Lav04} 
893: ($R/h\simeq 10^{-1}$, $K/(\gamma R)\simeq 10^{-4}$)
894: we find $V_{\rm men} \simeq 10^{-14}\, k_B T\;\ln(R/d)$, which is 
895: unimportant for the actual intercolloidal interaction. 
896: %
897: %{\tt I have corrected the numerical estimate of the amplitude, which
898: %  is now much smaller than quoted before. I have done the same in the
899: %  next subsection, and in the Conclusions with the power-law
900: %  dependence on $R/h$. It seems that you inadvertendly quoted the
901: %  amplitude of the \mbox{{\em deformation} ($\propto \Delta F$)}
902: %  whenever you really meant the amplitude of the \mbox{{\em
903: %      interaction energy} ($\propto (\Delta F)^2$)}. Please check.}
904:  
905: The elastic potential $V_{\rm el}(d)$ between two colloids 
906: is the interaction between the second quadrupole 
907: and the first quadrupole together with its string of image quadrupoles.
908: Using the solution given in Refs.~\cite{Ram96,Ruh97} we have checked
909: that $V_{\rm el}(d)$ remains repulsive.  For $d < h$ the overall magnitude of $V_{\rm el}(d)$
910: is somewhat weakened,
911: whereas for $d \gg h$ a crossover to $V_{\rm el}(d) \propto \exp(-d/h)$ is observed\footnote{This result can be obtained more easily by solving the field equations 
912: $\Delta n_i=0$ in cylindrical coordinates rather than by using the image quadrupoles.}.   
913: 
914: 
915: \subsubsection{Parallel anchoring at the bottom substrate }
916: 
917: We assume that the substrate induces a preferred in--plane axis for the director 
918: orientation which we take to be the  $x$--axis. With no colloid present at the nematic--air 
919: interface, the equilibrium director field is given by
920: \bea
921:  \label{eq:efield}
922:   \vect n_0=  \begin{pmatrix}  \sin(-q_0 z) \\ 0 \\ \cos(-q_0 z) \end{pmatrix}
923:   \;, \qquad q_0 = \pi /(2 h) \;,
924: \eea 
925: with the consequence that both at the nematic--air and at the nematic--substrate 
926: interface a constant stress is acting:
927: \bea
928:   \pi_{0,zz} = \frac{K}{2}\,q_0^2\;.
929: \eea
930: For the unperturbed interface to be in equilibrium, the air and substrate pressures
931: differ from the pressure in the liquid crystal: $p-p_{\rm sub\,[air]}=
932: \pi_{0,zz}$. 
933: 
934: We now introduce a single colloid at the nematic--air interface in the reference 
935: configuration. 
936: Due to this pressure difference, the force on the colloid
937: and the integrated stress over the nematic--air interface are
938: given  by (see Eq.~(\ref{eq:Ftot1}))
939: \bea
940:  \label{eq:f1}
941:   F\;{\bf e}_z &=& \int_{S_1} d{\bf A}\cdot ({\boldsymbol \pi}-\pi_{0,zz}\,{\bf 1})\;, \\
942:  \label{eq:fpi1}
943:   F_\pi &=& \int_{S_{\rm men}}  dA \; (\pi_{zz}- \pi_{0,zz}) \; ,
944: \eea
945: and the total excess force on the system ``colloid  and nematic--air interface"
946: is determined by
947: \bea
948:  \label{eq:df2}
949:   F- F_\pi &=& \Delta F = - \int_{S_{\rm sub}}  dA \; (\pi_{zz}-\pi_{0,zz})\;.
950: \eea
951: In order to calculate the director field $\vect n$ and the stress tensor ${\boldsymbol\pi}$
952: in the presence of the colloid, we introduce the auxiliary director deformation 
953: fields $v(x,y,z)$ and $w(x,y,z)$ which parametrize the deviations from the unperturbed 
954: director field $\vect n_0$ and which are small at large distances from the colloid:
955: \bea
956:  \label{eq:nlin}
957:   \vect n &=& \begin{pmatrix} \sin(-q_0 z + v)\cos w \\ \sin w \\ \cos(-q_0 z + v)\cos w
958:             \end{pmatrix} \approx \vect n_0 + \\
959:  \nonumber  & &
960:    \begin{pmatrix}  \cos(q_0 z)\;v + \frac{1}{2}\sin(q_0 z)(v^2+w^2) \\ w \\ \sin(q_0 z) \;v -\frac{1}{2}\cos(q_0 z)(v^2+w^2)  \end{pmatrix}
961:   + O((v,w)^3) \;.\quad
962: \eea
963: The first equality in Eq.~(\ref{eq:nlin}) is a general parametrization of
964: the director field $\vect n$ in terms of the auxiliary fields $v,w$ 
965: which fulfills $\vect n^2=1$.
966: These auxiliary fields are taken to be zero at the substrate and the nematic--air interface,
967: i.e., the boundary conditions are $v(x,y,0)=v(x,y,-h)=w(x,y,0)=w(x,y,-h)=0$.
968: The nematic free energy of the film up to order $O(v^2,w^2)$ is obtained
969: by inserting Eq.~(\ref{eq:nlin}) into Eq.~(\ref{eq:f_lc}) for the Frank  free energy
970: after dropping the total divergence of the $K_{24}$--type.
971: Using the boundary conditions for $v$ and $w$ we obtain \cite{Fuk02}
972: (with the notation introduced in Eq.~(\ref{eq:stresstensor})):
973: \bea
974:   {\cal F}^{\rm film}_{\rm ne} = {\cal F}_0 + \frac{K}{2}\int_{V_{\rm film}} d^3r 
975:     \left( v_{,i}^2 + w_{,i}^2 - q_0^2\;w^2 \right)\;.
976: \eea
977: Here, ${\cal F}_0$ is the free energy of the film without colloid. At first
978: glance it is not evident  that
979: this free energy is positive definite. However, the boundary conditions
980: on $v$ and $w$ ensure positivity \cite{Fuk02}. Upon minimization we find
981: $\Delta v=0$, i.e., the deformation field  $v$ of the director 
982: can again be expanded in terms of electrostatic multipoles (Eq.~(\ref{eq:es_mul})).
983: On the other hand, the solution for $w$ must fulfill the Helmholtz equation
984: $(\Delta + q_0^2)w=0$ and can be expanded in terms of multipoles as follows:
985: \bea
986:  \label{eq:hel_mul}
987:   w(r,\theta,\phi)& =& \frac{1}{\sqrt{q_0 r}}\; \sum_{j=0}^\infty\sum_{m=-j}^{j} 
988:   Y_{jm}(\theta,\phi)\; \\
989:   \nonumber  & & \left( w^J_{jm}\; J_{j+1/2}(q_0 r) + w^Y_{jm}\; Y_{j+1/2}(q_0 r) \right)\;
990:     ,
991: \eea
992: where $J[Y]_{j+1/2}(r)$ are the spherical Bessel functions of the first
993: [second] kind and $Y_{jm}(\theta,\phi)$ 
994: are the usual spherical harmonics for the standard set of spherical coordinates
995: $r,\theta,\phi$. 
996: (The origin is again taken as the center of the circular three--phase contact line.)
997: The coefficients $w^{J[Y]}_{jm}$ are dimensionless multipole moments.
998: 
999: The Dirichlet boundary conditions for $v$ and $w$ at the two interfaces can be fulfilled
1000: as before by constructing the full solution in terms of multipoles around the colloid
1001: and corresponding image multipoles as shown in Fig.~\ref{fig:ref1}. 
1002: Since rotational covariance is broken by the parallel substrate anchoring,
1003: the solution for $v$ contains a nonzero dipole contribution.
1004: Nevertheless, the director field
1005: still obeys a reflection symmetry
1006: with respect to the $xz$--plane: $v(x,y,z)=v(x,-y,z)$ and
1007: $w(x,y,z)=-w(x,-y,z)$. Therefore, the
1008: leading asymptotic behavior for $v$  is given by
1009: \bea
1010:  \label{eq:uasy}
1011:   v = P_v \frac{z}{r^3} + Q_v \frac{zx}{r^5} +\dots + v_{\rm image} \;.
1012: \eea
1013: The dipole contribution should vanish for $h\to \infty$. If one
1014: assumes a power--law dependence on $h$, dimensional analysis for 
1015: $P_v$ leads to 
1016: \bea
1017:  \label{eq:kappadef}
1018:  P_v = O(R^2\;(R/h)^\kappa) \qquad (\mbox{with}\; \kappa>0)\;.
1019: \eea
1020: The precise functional form of $P_v$  turns out to be unimportant for 
1021: the subsequent calculations.
1022: We note that an asymptotic solution with a nonvanishing
1023: $x$--component of the dipole moment cannot appear because it would not fulfill
1024: the boundary conditions and the reflection symmetry $w(x,y,z)=-w(x,-y,z)$ excludes
1025: any dipolar contribution in the solution for $w$.\footnote{
1026: In this respect the pictorial argument given in Ref.~\cite{Lav04} (see Fig.~3(b) therein)
1027: is slightly misleading (at least in an asymptotic sense): there the
1028:  assumed director configuration around the colloids for the
1029: case of a finite film thickness is a tilted dipole with a nonvanishing component
1030: $v \propto P_{v,x} x/z^3$. 
1031: Actually, the broken symmetry in $x$--direction rather
1032: enters through a tilted quadrupole.}  Therefore the leading asymptotic
1033: behavior of $w$ takes the form
1034: \bea
1035:   w &=&  \frac{zy}{r^2}\;\frac{1}{\sqrt{q_0 r}} \left( Q^J_w\;J_{5/2}(q_0 r) +
1036:    Q^Y_w \;Y_{5/2}(q_0 r) \right) + \dots 
1037:   \nonumber \\&& + w_{\rm image} \;.
1038:  \label{eq:vasy1}
1039: \eea
1040: We note that there are two quadrupolar contributions (with dimensionless moments
1041: $Q^J_w$ and $Q^Y_w$)
1042:  which show an oscillatory behavior 
1043: for radial distances $r \gg h$ (i.e., $q_0 r \gg 1$). On the other hand, for 
1044: large film thicknesses
1045: there is an intermediate asymptotic regime $R \ll r \ll h$:
1046: \bea
1047:  \label{eq:vasy2}
1048:   w (q_0 R \ll q_0 r \ll 1) &=& \frac{zy}{r^2}\;\sqrt{\frac{2}{\pi}} \left(
1049:     Q^J_w\;\frac{(q_0 r)^2}{15} -   Q^Y_w \; \frac{3}{(q_0 r)^3} \right)
1050:  \nonumber \\
1051:   && + \dots + 
1052:    w_{\rm image} \;.
1053: \eea
1054: For large film thicknesses the solution for the director field
1055: near the nematic--air interface in the regime $R \ll r \ll h$ should coincide
1056: with the solution for macroscopically thick films (Eq.~(\ref{eq:quadrupole})).
1057: Near the nematic--air interface, one has $v\approx n_x$, $w\approx n_y$. Therefore
1058: one recovers the rotationally covariant quadrupole solution of Eq.~(\ref{eq:quadrupole})
1059: for  $Q_v=Q$ (obtained by equating Eq.~(\ref{eq:uasy}) with Eq.~(\ref{eq:quadrupole})
1060: for $i=1$, i.e., $r_1=x$)  and $-3\sqrt{\di 2/\pi}\,Q^Y_w/q_0^3=Q$ (obtained by equating 
1061: Eq.~(\ref{eq:uasy}) with Eq.~(\ref{eq:quadrupole}) for $i=2$, i.e., $r_2=y$).
1062: Since $Q^Y_w \sim q_0^3 Q$ and $Q = O(R^3)$, $Q^Y_w=O([q_0 R]^3)$ is
1063: a very small number.
1064: 
1065: The Dirichlet boundary conditions for $w$ at the substrate and at the 
1066: nematic--air interface
1067: enforce that the contribution to $w$ due to the quadrupole $Q^J_w$ and all corresponding image
1068: quadrupoles is zero. This holds also for the contribution of  all higher multipoles of 
1069: degree $j$ 
1070: for which the  radial dependence is given by $J_{j+1/2}(r)$. 
1071: The only solution of the Helmholtz equation
1072: which fulfills the boundary conditions and which, as an additional requirement, 
1073: is smooth everywhere,
1074: is $w\equiv 0$. Since the Bessel functions of the first kind are smooth
1075: everywhere, all respective multipole moments must be zero.
1076: (This does not hold for the multipole moments pertaining to the Bessel functions
1077: of the second kind since $Y_{j+1/2}(r)$ is singular at $r=0$.) 
1078: % Note that this is valid
1079: %even for more general boundary conditions at the substrate derived from
1080: %Eq.~(\ref{eq:fs_lc}).
1081: 
1082: The excess force on the system ``colloid and nematic--air interface" follows from
1083: Eq.~(\ref{eq:df2}) and can be expressed as
1084: \bea
1085:  \label{eq:stress_twisted}
1086:  \Delta F = - \frac{K}{2}\int_{S_{\rm sub}} dA\; (v_{,z}^2-2q_0\,v_{,z}+w_{,z}^2) \;,
1087: \eea
1088: utilizing the boundary conditions %and symmetries of 
1089: for the solutions $v$ and $w$.
1090: For the multipoles appearing in the expansion of $v$ the method described in 
1091: Subsec.~\ref{sec:aligned}
1092: may be used. In order to obtain  the quadrupolar contributions to $w$ we perform the summation over the
1093: image multipoles and the integration over the substrate surface numerically. Due to the
1094: slow decay of $Y_{5/2}(r) \propto (\cos r)/r^{1/2}$ the stress integral is superficially
1095: divergent. However, a detailed asymptotic analysis yields convergence with the result  
1096: \bea
1097:  -\Delta F = \pi \,K \left( \frac{3\zeta(3)}{2}\,\frac{P_v^2}{h^4} + \frac{5\zeta(5)}{12}\,
1098:   \frac{Q_v^2}{h^6} + c_w\,(Q^Y_w)^2 \right) \;,\quad
1099: \eea
1100: where $c_w\approx 0.30$ has been determined numerically. 
1101: %{\tt I have rewritten the following sentence:} 
1102: Note that the linear contribution
1103: %s in $P_v$ and $Q_v$ 
1104: due to the term $\sim q_0\,v_{,z}$
1105: in the stress tensor (Eq.~(\ref{eq:stress_twisted})) turns out to be zero, as can be 
1106: easily checked by applying Gauss' theorem and the field equation $\Delta v = 0$.
1107: Using the above considerations
1108: concerning the magnitude of the multipole moments $P_v$, $Q_v$ and $Q^Y_w$ we obtain
1109: \bea
1110:  \label{eq:ftwist}
1111:  -\frac{\Delta F}{K} \sim a_P \,\left(\frac{R}{h}\right)^{4+2\kappa} + 
1112:   a_Q\, \left(\frac{R}{h}\right)^6\;,
1113: \eea
1114: where $a_P$ and $a_Q$ are numerical coefficients of order 1 and $\kappa$ has been
1115: defined in Eq.~(\ref{eq:kappadef}). Again the excess force
1116: on the colloid falls off rapidly with increasing film thickness such that the induced 
1117: logarithmic capillary interaction (see Eq.~(\ref{eq:vmen_noniso})) remains very weak.
1118: For the parameters characterizing the experimental system studied in Ref.~\cite{Lav04} 
1119: ($R/h\sim 10^{-1}$, $K/(\gamma R)\sim 10^{-4}$)
1120: we find $V_{\rm men}(d) \sim$ \linebreak $10^{-11-4\kappa}\, k_B T\;\ln(R/d)$ 
1121: which appears to be undetectably small.
1122: Note that for $d<h$ the direct elastic repulsion remains essentially unchanged
1123: because in this regime the leading term of the elastic interaction is given by the
1124: repulsion between the two quadrupoles located at the colloid sites. In this case the image
1125: quadrupoles can be neglected. 
1126: 
1127: 
1128: 
1129: 
1130: \section{Discussion and conclusion}
1131: \label{sec:conclusion}
1132: 
1133: We have investigated the effective potential between two colloidal microspheres
1134: of radius $R$
1135: floating at asymptotically large distances $d$ on an interface between a nematic film 
1136: of thickness $h$ and air (Fig.~\ref{fig:nem_exp}).
1137: This effective potential is the sum of an elastic interaction caused by
1138: the director distortions around the colloids and of a capillary interaction mediated
1139: by surface deformations. We have analyzed the effective potential for large $d$ employing the
1140: coarse--grained Frank free energy (within the one constant approximation)
1141: for the director distortions and the
1142: linearized Young--Laplace equation for the interface distortions.   
1143: 
1144: In the case of a macroscopically thick nematic film, the director deformation
1145: around a single colloid is of quadrupolar type. Thus the induced elastic interaction
1146: between two colloids at distance $d$ is repulsive and of quadrupolar type  $\propto d^{-5}$.
1147: The capillary interaction is also repulsive and decays $\propto d^{-5}$ but it is much
1148: weaker than the elastic interaction. This rapid decay is a consequence of the mechanical 
1149: isolation of the system,
1150: %
1151: i.e., of the fact that the net force on the colloidal particles and
1152: the surrounding interface vanishes.
1153: 
1154: A finite film thickness $h$ leads to a violation of the mechanical isolation of the
1155: system ``colloid and interface". (The excess force is counteracted by the film substrate
1156: such that the whole experimental system is of course in mechanical equilibrium.)
1157: However, on the thermal energy scale the strength of the ensuing logarithmic 
1158: capillary potential turns out to be very small. 
1159: In the case of homeotropic boundary conditions on both sides of the nematic film
1160: the strength is proportional to $(R/h)^{12}$ (see Eq.~(\ref{eq:fqq}) and the subsequent discussion).
1161:  For  twisted boundary conditions
1162: (parallel at the bottom substrate, perpendicular at the upper nematic--air interface)
1163: there is a qualitatively different asymptotic behavior of the director field
1164: due to a dipole contribution, which is induced by the broken rotational 
1165: symmetry and vanishes for $h\to \infty$.
1166: However, even for these boundary conditions 
1167: %{\tt Here I have dropped the following sentence: ``since this dipole contribution induced by the finite film thickness
1168: %must vanish for $h\to \infty$'', because the amplitude of the log would also vanish even if the dipole would not (Eq.~(43) with $\kappa=0$).}
1169: the strength of the logarithmically varying 
1170: capillary potential remains 
1171: extremely
1172: small, vanishing at least $ \propto (R/h)^8$ (see Eq.~(\ref{eq:ftwist}) and the subsequent discussion).
1173: % and thus seemingly does not offer an explanation of 
1174: %the recent experimental observations made in Ref.~\cite{Lav04}.
1175: 
1176: Thus our analysis based on the mechanical isolation of the
1177: experimental system under consideration rules
1178: out a significant attractive contribution $\sim \ln d$ of capillary type 
1179: to the  effective potential
1180: between two colloids at a nematic interface.
1181: The amplitude of such a logarithmic contribution vanishes rapidly 
1182: for large  film thickness $h$. Therefore, this effective pair potential 
1183: does not provide a mechanism for the stability of isolated colloid clusters
1184: at a nematic interface as reported in Ref.~\cite{Lav04}.
1185: %We have demonstrated that the explanation offered in Ref.~\cite{Lav04}
1186: %for the , based on an effective 
1187: %pair potential between
1188: %the colloids, is not valid.   Also the elastic contribution of dipolar type
1189: %($\propto d^{-3}$) to the effective pair potential is not backed by our asymptotic 
1190: %analysis.   
1191: %Ref.~\cite{Lav04}
1192: %derives an $h$--independent amplitude because only part of the forces
1193: %contributing to the net mechanical balance are identified. 
1194: 
1195: 
1196: Naturally one strives for other explanations for the observation of stable clusters
1197: reported in Ref.~\cite{Lav04}. 
1198: There the mutual center--to--center distance between 
1199: neighboring colloids in the cluster has been found to be between $3R$ and $5R$, 
1200: depending on the
1201: radius $R$ of the colloids. If the cluster stability is attributed to a minimum
1202: in the effective pair potential at this range $3R \dots 5R$ of distances, 
1203: the applicability of the above asymptotic considerations at such distances is doubtful 
1204: for both the elastic and capillary
1205: contributions to the effective pair potential: 
1206: \begin{itemize}
1207:  \item {\em Elastic part:} For the single--colloid problem, the asymptotic, quadrupolar
1208:   form of the director field (Eq.~(\ref{eq:quadrupole})) is based on the assumption
1209:   that the deviations of the director from the preferred direction $\vect e_z$ are small: \linebreak
1210:   $|n_{x[y]}| \ll 1$. At a radial distance $\rho$ from the center
1211:   of the colloid this implies $Q/\rho^3 \ll 1$, which seems to be fulfilled for
1212:   the dimensional estimate for the quadrupole moment $Q = O(R^3)$ and for the
1213:   distances $\rho = 3R \dots 5R$ under discussion. However, the absolute magnitude of the
1214:   director deformations is not fixed by these arguments, so that it might be
1215:   worthwhile to determine the minimum of the Frank free energy 
1216:   by a full numerical calculation in the presence of one or two colloids at
1217:   the nematic--air interface, similar to Ref.~\cite{Sta99} where  the explicit 
1218:   director solution around a single colloid in the bulk has been determined.   
1219:   Our analysis indicates that 
1220:   only if the multipolar expansion of the director field around a single trapped colloid
1221:   fails for distances $\rho = 3R \dots 5R$, there is a chance for attractive elastic
1222:   interactions between two colloids at these distances. 
1223:   Failure of the multipole expansion at these intermediate 
1224:   distances would point to a strong deviation of the magnitude of the 
1225:   multipole moments $M_n$ of order $n$ from the dimensional estimate
1226:   $M_n \sim R^{n+1}$. Thus it is certainly worthwhile to determine
1227:   multipole renormalizations for bulk and interface configurations, taking into
1228:   account the full nonlinearity of the director field equations.
1229: 
1230:  \item {\em Capillary part:} Whenever the colloid--induced nematic stress deviates from the
1231:   asymptotic, quadrupolar form (Eq.~(\ref{eq:pizz_single})), corrections
1232:   to the asymptotic capillary potential will
1233:   arise. However, due to the smallness of the dimensionless
1234:   force $\varepsilon_F$ (see Eq.~(\ref{eq:epsdef}), $\varepsilon_F \propto K/(\gamma R) \approx 10^{-3}$) 
1235:   the perturbative treatment of the ensuing capillary deformations
1236:   is justified. Therefore, one of our main findings, i.e., that the magnitude
1237:   of the capillary potential ($\propto \varepsilon_F^2$) is always smaller
1238:   than the magnitude of the elastic interaction ($\propto |\varepsilon_F|$), 
1239:   is likely to hold also for elastic stresses deviating from the
1240:   asymptotic limit.
1241: %{\tt In case we decide to answer also the claim with the torques by
1242: %  Lavrentovich et al.}: 
1243: 
1244: Another possible, subtle effect related to the 
1245: anisotropy inherent to nematic phases
1246: allows the substrate to exert also a torque on the colloidal particles
1247: and the surrounding interface. This case can be studied similarly
1248: to what we have done here. Actually, the result can be obtained 
1249: straightforwardly if the analogy of capillary deformation with 2D
1250: electrostatics is employed (see, e.g., Ref.~\cite{Dom06}): The net
1251: force creates ``capillary monopoles'' $Q_{\rm cap}$ with an interaction
1252: energy $V_{\rm men}(d) = Q_{\rm cap}^2/(2\pi\gamma)\, \ln( C\lambda/d)$ 
1253: (see Eq.~(\ref{eq:vmen_noniso})). A net torque creates
1254: ``capillary dipoles'' ${\bf P}_{\rm cap}$ and the interaction energy
1255: between them is $V_{\rm men}(d) \sim (|{\bf P}_{\rm cap}|/d)^2$. This decay is
1256: sufficiently
1257: slow to dominate asymptotically the
1258: nematic--mediated repulsion. However, dimensional analysis yields $P_{\rm cap}
1259: \sim R\, Q_{\rm cap}$ so that the amplitude of
1260: this capillary energy, which must vanish for large film thicknesses $h\to\infty$, 
1261: has also a
1262: too small numerical value to explain the experimental results reported in Ref.~\cite{Lav04}.
1263: \end{itemize}
1264: %
1265: %First, we want to emphasize that our strictly asymptotic
1266: %considerations cannot exclude the possibility of stronger,
1267: %non--logarithmical capillary attractions when the colloids are close
1268: %to each other.
1269: %
1270: %Actually, we obtain indications that beyond the small--deformation
1271: %regime of validity of our calculations, a capillary attraction might
1272: %arise which would dominate asymptotically over the nematic--mediated
1273: %repulsion. There is, however, no evidence that the experiments
1274: %conducted in Ref.~\cite{Lav04} involve large interfacial deformations.
1275: 
1276: We recall that for simplicity our calculations have been carried out in the ``strong
1277: anchoring'' limit. ``Weak anchoring'' would lead to a
1278: different director field, and thus to different values of the
1279: dimensionless forces $\varepsilon_F$ and $\varepsilon_\pi$ (see
1280: Eq.~(\ref{eq:epsdef})). It would also lead to an explicit contribution to the
1281: expression~(\ref{eq:F}) for the
1282: free energy accounting for the anchoring ``wetting''
1283: energies~(Eq.~(\ref{eq:fs_lc})), which in the ``strong
1284: anchoring'' limit considered so far were subsumed in a (quantitatively
1285: negligible) additive renormalization of the surface tensions. The consequences of 
1286: this more general case of no strong anchoring have to
1287: be explored yet, but we do not expect that this alters our
1288: conclusion that a logarithmically varying capillary attraction is
1289: ruled out by mechanical isolation ($\varepsilon_F = \varepsilon_\pi$),
1290: because this latter result is based on very general principles.
1291: 
1292: 
1293: Finally we note that for a two--dimensional system, the interactions between 
1294: two upright circular colloids trapped at the nematic--isotropic interface have been
1295: studied numerically \cite{Tas05} using a Landau--de Gennes free energy. 
1296: In three dimensions, this corresponds to
1297: long, cylindrical colloids which are aligned side--to--side at the interface.
1298: For boundary conditions that yield a similar, quadrupolar behavior of the director 
1299: field around a single colloid \cite{Pet98}, 
1300: the effective interaction is found to be repulsive
1301: and consistent with a power--law decay (for intermediate distance $d/R$ up to 7). 
1302: In this particular two--dimensional situation, the repulsive interactions appear
1303: to be longer--ranged $\propto d^{-1}$ and in the numerical results 
1304: for the director field there is no trace of a sizeable interface
1305: deformation which would lead to capillary attractions.
1306: 
1307: In summary, we have presented several arguments 
1308: % that point to the
1309: % absence of an effective potential between two colloids at a nematic interface
1310: % which consists of a repulsive elastic and an attractive capillary part. 
1311: which rule out an asymptotic attraction of capillary origin in the
1312: effective interaction potential between two colloids at a nematic interface.
1313: If one discards effective pair potentials as the source of the
1314: stability of colloidal mesostructures as the ones found in Ref.~\cite{Lav04}, 
1315: the question arises whether this stability is a consequence
1316: of genuine many--body effects \cite{Lav07a}. Nematic surfaces (without colloids)
1317: can stabilize regular patterns of surface defects \cite{deG74}. If these  defects
1318: persist in the presence of colloids, the experimentally observed regular order of 
1319: the colloids might
1320: be attributed to them. This issue calls for further experimental and theoretical
1321: investigations.   
1322: 
1323: 
1324: 
1325: {\bf Acknowledgment:}
1326: We acknowledge useful discussions with S.~Chernyshuk, O.~Lavrentovich,
1327: B.~Lev, P.~Patr\'icio, and J.~M.~Romero--Enrique. M.~O. thanks the German 
1328: Science Foundation (DFG) for financial
1329: support through the Collaborative Research Centre (SFB-TR6) 
1330: ``Colloids in External Fields", project section D6-NWG.
1331: %
1332: A.~D. acknowledges financial support from the Junta de Andaluc{\'\i}a
1333: (Spain) through the program ``Retorno de Investigadores''.
1334: 
1335: 
1336: 
1337: 
1338: 
1339: 
1340: \begin{appendix}
1341: 
1342: \section{Corrections to the strong anchoring limit}
1343: 
1344: \label{sec:app1}
1345: 
1346: Here we consider the asymptotic director field around a single colloid at the nematic--air
1347: interface in the case of finite anchoring strength. For large distances from the colloid,
1348: the overall deviations from the preferred director $\vect n =(n_1,n_2,n_3) 
1349: = (0,0,1)$ are small
1350: and the total free energy of the nematic phase is given by the harmonic approximation
1351: for the bulk part (Eq.~(\ref{eq:fasy})) and the Poulini expression for the surface part:
1352: \bea
1353:   {\cal F}_{\rm ne} & = & \frac{K}{2} \spaceint{V_{\rm ne}}{r} 
1354:      \sum_{i=1,2} \left( \nabla n_i \right)^2 +
1355:      \frac{W_1}{2} \int_{A_{\rm air-ne}}\!\!\! dA \;
1356:    (\vect n \cdot \vect e_A)^2 \nonumber  \\
1357:    & =& \frac{K}{2} \spaceint{V_{\rm ne}}{r} \left( \left( \nabla n_1 \right)^2 +
1358:   \left( \nabla n_2 \right)^2 \right) + 
1359:  \nonumber \\ &&  
1360:   \frac{W_1}{2} \int_{A_{\rm air-ne}}\!\!\! dA \;
1361:   (1-n_1^2-n_2^2)  \;.
1362: \eea   
1363: The second equation holds because the normal $\vect e_A$ of the nematic--air interface
1364: $A_{\rm air-ne}$ is parallel to the $z$--direction and $\vect n^2=1$. 
1365: For particles with radii $R$ in the $\mu$m range, such as those investigated 
1366: in Ref.~\cite{Lav04}, 
1367: the dimensionless parameter $\alpha = K/(|W_1| R)$ is smaller than 1. Minimizing
1368: the free energy renders the Laplace equation in the bulk,
1369: \bea
1370:    \Delta n_i = 0 \qquad (i=1,2)\,,
1371: \eea 
1372: supplemented with the Robin boundary condition:
1373: \bea
1374: %     K \frac{\partial}{\partial z} n_i (x,y,z=0) - W_1 n_i (x,y,z=0) & = & 0 
1375: %    \quad \to \nonumber \\
1376:    \alpha R\, \frac{\partial}{\partial z} n_i (x,y,z=0) - n_i (x,y,z=0) & = & 0
1377:   \quad (i=1,2) \;. \nonumber \\
1378:  \label{eq:boun}
1379: \eea
1380: The general, asymptotic solution  is given by the multipole {\em ansatz} 
1381: (Eq.~(\ref{eq:es_mul}))
1382: \bea
1383:   n_i &=& q_i\,\frac{1}{r} + \sum_{\beta=1}^3 P_{i\beta}\,\frac{r_\beta}{r^3} +  
1384:       \sum_{\beta,\gamma=1}^3 Q_{i\beta\gamma}\,\frac{r_\beta\,r_\gamma}{r^5}
1385:  + \dots
1386: \eea
1387: with the requirement of rotational covariance around the $z$--axis. In view of
1388: the boundary condition (Eq.~(\ref{eq:boun})) we note that the derivative
1389: in Eq.~(\ref{eq:boun}) applied to a multipole of order $m$ generates 
1390: a multipole term of order
1391: $m+1$. Thus, to leading order in $1/r$, the boundary condition can only be met if
1392: the leading multipole ($n_i^{\ell}$) fulfills the strong anchoring condition
1393: $n_i^{\ell}(z=0)=0$ and is accompanied by a subleading multipole ($n_i^{s\ell}$)
1394: which is connected to the leading multipole through the condition
1395: $\alpha R\, \partial_z n_i^{\ell}(z=0) = n_i^{s\ell} (z=0)$.
1396: Both the monopole and the dipole do not fulfill the strong anchoring condition
1397: $n_i^{\ell}(z=0)=0$.
1398: Therefore, as before, we find that the multipole expansion starts with the quadrupole solution
1399: and that it is necessarily accompanied by a hexapole:
1400: \bea
1401:  \label{eq:app_sol1}
1402:   n_i & = & n_i^{\ell} + n_i^{s\ell} \;, \\
1403:   n_i^{\ell} &=& Q\,\frac{z r_i}{r^5} + Q^{\rm mag} \epsilon_{i\alpha 3}\, \frac{z r_\alpha}{r^5}\;, \\
1404: %  \mbox{and} \quad 
1405:   n_i^{s\ell} &=& \alpha QR\, \frac{r_i(r^2-5z^2)}{r^7} +
1406:                \alpha Q^{\rm mag}R \epsilon_{i\alpha 3}\, \frac{r_\alpha(r^2-5z^2)}{r^7}
1407:   \;. \nonumber \\
1408:  \label{eq:app_sol2}
1409: \eea
1410: Since $\partial_z n_i^{s\ell}(z=0)=0$, this solution 
1411: (Eqs.~(\ref{eq:app_sol1})--(\ref{eq:app_sol2})) exactly satisfies the boundary condition in 
1412: Eq.~(\ref{eq:boun}). Thus
1413: the asymptotically dominant director field consists of a quadrupole--hexapole
1414: superposition. 
1415: Furthermore, the magnitude of the accompanying hexapole moment is small due to
1416: the factor $\alpha$, therefore the results reported before for strong anchoring 
1417: (based on the
1418: leading quadrupole only) are unaffected by a finite anchoring strength at the nematic--air
1419: interface.   
1420: 
1421: 
1422: 
1423: \section{Force balance in a general configuration} 
1424: 
1425: \label{sec:app2}
1426: 
1427: \begin{figure}[t]
1428:  \begin{center}
1429:   \epsfig{file=forcebalance.eps, width=\columnwidth}
1430:  \end{center}
1431:  \caption{ 
1432:    \label{fig:forcebalance}
1433:    A general configuration of the particle and the fluid interface.
1434:    The arrows indicate the orientation of the corresponding surfaces:
1435:    $S_{\rm men}$ is the interface, $S^{1(2)}_{\rm part}$ is the
1436:    surface of the particle in contact with the lower (upper) fluid
1437:    phase, $S_{\rm wall}$ is the surface of the container of the system
1438:    (sketched here as a quadrangular box for simplicity). The
1439:    three--phase contact line between the particle and the interface is
1440:    denoted as $C_0$.}
1441: \end{figure}
1442: 
1443: For the benefit of the reader, we discuss in this Appendix some
1444: previous results \cite{Dom05,Dom06} concerning the  force balance of a general
1445: equilibrium configuration and demonstrate how the amplitude of an 
1446: asymptotic, logarithmically varying interfacial deformation is
1447: determined solely by this mechanical condition of force balance.
1448: 
1449: Figure~\ref{fig:forcebalance} represents a colloidal particle in equilibrium at
1450: the deformed interface; in general the deformation is not small. 
1451: The condition of mechanical equilibrium implies that locally
1452: %not only that the net force on the whole system , but also 
1453: the net force on any part of the system must vanish. Thus one
1454: has:
1455: \begin{enumerate}
1456: \item Each of the fluid phases is in equilibrium. 
1457: %The lower (nematic) phase 
1458: %   , so that the integral of the
1459: %   stress ${\mathbf \Pi}$ over the surface enclosing the phase must
1460: %   vanish:
1461: %   \begin{equation}
1462: %     \int_{S_{\rm men} \cup S^{\rm wall}_{1} \cup (\mbox{}-S_1)}
1463: %     d{\bf A} \cdot {\mathbf \Pi} = {\bf 0} .
1464: %   \end{equation}
1465:   We introduce the %(elastic and hydrostatic)
1466:   forces exerted {\em by} each fluid phase  on the particle, on the whole
1467:   fluid meniscus, and on the wall of the container as
1468:   %
1469:   \bea
1470:   \label{eq:fpart}
1471:   {\mathbf F}_{\rm part}^{1(2)} & := & \int_{S_{\rm part}^{1(2)}} d{\bf A} \cdot {\mathbf \Pi}^{1(2)} , \\
1472:   \label{eq:fint}
1473:   {\mathbf F}_{\rm men}^{1(2)} & := & \mbox{} -(+) \int_{S_{\rm men}} d{\bf A} \cdot {\mathbf \Pi}^{1(2)} , \\
1474:   {\mathbf F}_{\rm wall}^{1(2)} & := & \mbox{} - \int_{S_{\rm wall}^{1(2)}} d{\bf A} \cdot {\mathbf \Pi}^{1(2)} ,  
1475:   \eea
1476:   % 
1477:   respectively, in terms of the stress tensor ${\bf \Pi}^{1(2)}({\bf r})$ in each
1478:   fluid phase with due account for the orientation of the surfaces
1479:   (see Fig.~\ref{fig:forcebalance}). The superscript $^{1(2)}$
1480:   indicates the lower (upper) phase in Fig.~\ref{fig:forcebalance}. 
1481:   The condition of mechanical equilibrium
1482:   of each phase under the influence of these three forces reads
1483:   \begin{equation}
1484:     \label{eq:netFfluid}
1485:     {\mathbf F}_{\rm part}^{1(2)} + {\mathbf F}_{\rm men}^{1(2)} +
1486:     {\mathbf F}_{\rm wall}^{1(2)} = {\bf 0} .
1487:   \end{equation}
1488: 
1489:   The total force exerted {\em by} the fluids on the particle is
1490:   ${\mathbf F}_{\rm part} := {\mathbf F}_{\rm part}^{1} + {\mathbf
1491:     F}_{\rm part}^{2}$, and on the meniscus it is ${\mathbf F}_{\rm
1492:     men} := {\mathbf F}_{\rm men}^{1} + {\mathbf F}_{\rm men}^{2}$.
1493:   The expressions for these forces reduce to those given in Eqs.~(\ref{eq:Fdef}) and
1494:   (\ref{eq:Fpi}), respectively, upon evaluation in the reference 
1495:    configuration depicted
1496:   in Fig.~\ref{fig:ref}.
1497:   %
1498: %   {\tt Note: strictly speaking, Eq.(\ref{eq:Fpi}) should be defined
1499: %     with the opposite sign}
1500:   
1501:   If the condition of mechanical equilibrium is applied 
1502:   locally to an infinitesimal volume in
1503:   the bulk of each of the fluid phases, it turns into $\nabla\cdot{\bf
1504:     \Pi}^{1(2)} = {\bf 0}$. In the nematic phase this condition yields
1505:   the field equations determining the director field.
1506: 
1507: \item The particle is in mechanical
1508:   equilibrium under the combined action of ${\mathbf F}_{\rm part}$
1509:   and the tension exerted {\em by} the interface on the particle at the
1510:   three--phase contact line $C_{0}$. This tension can be expressed in terms of a line
1511:   integral involving the surface tension $\gamma$:
1512:   \begin{equation}
1513:     {\mathbf F}_{\rm contact} := \mbox{} - \gamma \oint_{C_0} d\ell 
1514:     \,{\mathbf e}_c , %\times {\mathbf e}_n ,
1515:   \end{equation}
1516:   where ${\mathbf e}_c$ is the unit vector tangent to the interface,
1517:   normal to the contact line, and oriented towards the particle side. 
1518:   Therefore, the condition of mechanical equilibrium reads
1519:   \begin{equation}
1520:     \label{eq:netFpart}
1521:     {\mathbf F}_{\rm part} + {\mathbf F}_{\rm contact} = {\bf 0} .
1522:   \end{equation}    
1523: 
1524: \item Any piece $S_{\rm int} \subset S_{\rm men}$ of the fluid
1525:   interface is in mechanical equilibrium. The force on $S_{\rm int}$
1526:   exerted by the fluid phases and the tension exerted on this piece at
1527:   its boundary, $\partial S_{\rm int}$, are balanced:
1528:   \begin{equation}
1529:     \label{eq:netFint}
1530:     \int_{S_{\rm int}} d{\bf A} \cdot 
1531:     [{\mathbf \Pi}^2 - {\mathbf \Pi}^1]
1532:     + \gamma \oint_{\partial S_{\rm int}} d\ell \,{\mathbf e}_c 
1533:     = {\bf 0} ,
1534:   \end{equation}
1535:   with ${\mathbf e}_c$ oriented towards the exterior of $S_{\rm int}$.
1536:   If this expression is applied locally to an infinitesimal piece of
1537:   interface, it turns into an equation for the interfacial deformation
1538:   relating the mean curvature of the interface to the pressure jump
1539:   accross it. If, in addition, the interface deviates only slightly
1540:   from a flat interface (identified with the plane $z=0$), this equation reduces
1541:   in turn to the well--known equation~(\ref{eq:linearYLeq}) below
1542:   %$\gamma \nabla^2 u = \Pi_{zz}^1 - \Pi_{zz}^2$ 
1543:   for the local height $u(x,y)$ over the plane $z=0$.
1544: 
1545: \end{enumerate}
1546: 
1547: The net force balance of the whole system follows %of course 
1548: from the three separate balance conditions~(\ref{eq:netFfluid}),
1549: (\ref{eq:netFpart}), and (\ref{eq:netFint}): with $S_{\rm int} =
1550: S_{\rm men}$ (so that $\partial S_{\rm int} = C_0 \cup C_{\rm wall}$;
1551: $C_{\rm wall}$ is the three--phase contact line between phase 1, phase 2, and the
1552: container enclosing the system),
1553: one finds that, as expected, in equilibrium the net force of the outer environment on
1554: the system must vanish\footnote{The reasoning can be
1555:   easily generalized to the case that in addition to the
1556:   short--ranged influence of the wall there are also external fields
1557:   (gravity, electric force) acting on any part of the system.}:
1558: % on the fluid phases by the wall of the container
1559: % must be balanced by a tension at the contact line of the fluid
1560: % interface with the wall
1561: \begin{equation}
1562:   \label{eq:netFtotal}
1563:   {\mathbf F}_{\rm wall}^{1} + {\mathbf F}_{\rm wall}^{2} -
1564:   \gamma \oint_{C_{\rm wall}} d\ell \,{\mathbf e}_c = {\bf 0} ,
1565: \end{equation}
1566: where ${\bf e}_c$ points to the exterior of the system.
1567: 
1568: The condition of mechanical equilibrium can be applied advantageously
1569: to obtain a precise statement about the amplitude of an
1570: interfacial deformation $u(\rho,\phi)$ varying logarithmically with
1571: the lateral distance $\rho \gg R$ from the particle  with radius $R$. 
1572: Far away from the particle, interface deformations and their gradients are
1573: small, so that the linearized equation holds.
1574: %Since a logarithmic dependence
1575: %is a solution of the 2D Laplace equation, it can only arise provided
1576: %there is a region where the deformation from a flat interface is small
1577: %and accordingly the linearized equation holds.
1578: % and there
1579: % is no pressure force deforming the interface, so that $\gamma \nabla^2
1580: % u =0$. Physically, this requires a clear spatial separation between
1581: % the regions where the forces deforming the interface are applied, in
1582: % our case the neighbourhood of the particle and that one of the wall.
1583: Thus there is a distance $\xi$ beyond which the linear theory is applicable:
1584: \begin{equation}
1585:   \label{eq:linearYLeq}
1586:   \gamma \nabla^2 u (\vect r) = \Pi_{zz}^1 (\vect r) - \Pi_{zz}^2(\vect r)  \qquad
1587:   ({\bf r} \in S_{\rm ext}) \;,  %\xi < \rho ,
1588: \end{equation}
1589: where the piece of interface $S_{\rm ext}$ is enclosed by the
1590: circle $\rho=\xi$ and the line $C_{\rm wall}$.  The general solution 
1591: to this inhomogeneous Laplace equation can be written as
1592: \begin{eqnarray}
1593:   \label{eq:generalu}
1594:   u(\rho,\phi) & = & A_0 + B_0 \ln\frac{\xi}{\rho} +  \\ && 
1595:   \sum_{m=1}^{+\infty} \left[ A_m \left(\frac{\rho}{\xi}\right)^m 
1596:     + B_m \left(\frac{\xi}{\rho}\right)^m \right] \cos m(\phi-\phi_m) +
1597:    \nonumber\\
1598:   & &  \frac{1}{2\pi\gamma} \int_{S_{\rm ext}} \!\!\! d\phi' d\rho' \rho'\;
1599:   [\Pi_{zz}^1({\bf r}') - \Pi_{zz}^2({\bf r}')] \, 
1600:   \ln\frac{|{\bf r}-{\bf r}'|}{\rho}\; . \nonumber
1601: \end{eqnarray}
1602: The fixed values of the constants $A_m$, $B_m$, and $\phi_m$ are determined by
1603: the boundary conditions. This expression reduces to
1604: Eq.~(\ref{eq:sol1}) in the particular case of rotational symmetry and
1605: a wall located at infinity.
1606: 
1607: We can apply Eq.~(\ref{eq:netFint}) to the piece $S_{\rm int}=S_{\rm
1608:   men}\backslash S_{\rm ext}$ enclosed by the contact line $C_0$
1609: and the circle $\rho=\xi$. Invoking Eq.~(\ref{eq:netFpart}), one has
1610: \bea
1611:   {\bf F}_{\rm part} + \int_{S_{\rm int}} d{\bf A} \cdot [{\mathbf \Pi}^2 - {\mathbf \Pi}^1]
1612:   &=& \mbox{} - \gamma \oint_{\rho = \xi} d\ell \,{\mathbf e}_c
1613:   \\ \nonumber
1614:   &\approx& \mbox{} - \gamma \, {\bf e}_z \int_0^{2\pi} d\phi \; \rho 
1615:   \left.\frac{\partial u}{\partial \rho} \right|_{\rho=\xi} ,
1616: \eea
1617: where the last, approximate equality involves the leading order term in
1618: $\nabla u$ of the line integral. (This approximation is justified because the interface 
1619: deviates only slightly from a flat interface at the circle $\rho=\xi$.) 
1620: Evaluating this latter term for the general
1621: solution~(\ref{eq:generalu}) one finally finds that the amplitude of
1622: the logarithmic term in Eq.~(\ref{eq:generalu}) is proportional to the
1623: force exerted by the upper and the lower fluid on the particle and on the meniscus:
1624: \begin{equation}
1625:   \label{eq:amplitudelog}
1626:   B_0 = \frac{1}{2\pi\gamma} {\bf e}_z \cdot \left[ 
1627:     {\bf F}_{\rm part} + {\bf F}_{\rm men} \right] .
1628: \end{equation}
1629: In obtaining Eq.~(\ref{eq:amplitudelog}) we have used that
1630: in the region $S_{\rm ext}$ (where the interface is almost flat)
1631: one has to leading order
1632: \begin{equation}
1633:   \int_{S_{\rm ext}} d{\bf A} \cdot [{\mathbf \Pi}^2 - {\mathbf \Pi}^1]
1634:   \cdot {\bf e}_z
1635:   \approx 
1636:   \int_{S_{\rm ext}} d\phi\, d\rho\, \rho \;
1637:   [\Pi_{zz}^2({\bf r}) - \Pi_{zz}^1({\bf r})] .
1638: \end{equation}
1639: In conclusion, if the stress $\Pi_{zz}^1({\bf r}) - \Pi_{zz}^2({\bf
1640:   r})$ decays sufficiently fast with the distance $\rho$ from the
1641: particle, the solution in Eq.~(\ref{eq:generalu}) will be dominated
1642: asymptotically by the logarithm with an amplitude given by
1643: Eq.~(\ref{eq:amplitudelog}), provided this amplitude does not
1644: vanish \footnote{To which extent this asymptotic regime is actually observable in
1645:   a particular experimental realization depends on the precise
1646:   functional form of ${\bf \Pi}^{1(2)}({\bf r})$ and the values of the
1647:   constant parameters $A_m$, $B_m$, and $\phi_m$ entering the solution given by
1648:   Eq.~(\ref{eq:generalu}).}. As one can infer from
1649: Eq.~(\ref{eq:netFfluid}), this latter condition means physically that
1650: the walls exert a non-vanishing force: a
1651: logarithmic term can only arise if mechanical isolation of the system ``colloid + interface'' is violated.
1652: (Note that Eq.~(\ref{eq:amplitudelog}) follows actually from
1653: Eq.~(\ref{eq:netFtotal}) if the interfacial deformations are small.)
1654: 
1655: We emphasize the generality of this result: it only requires that the
1656: interface departs slightly from a flat one for distances sufficiently far
1657: from the particle. (Otherwise it is actually not useful to
1658: speak of a logarithmically varying deformation to begin with.) In
1659: particular, the interfacial deformation close to the particle may be
1660: arbitrarily large, the particle itself need not be perfectly spherical,
1661: or it may even consist of a many--body configuration lacking any kind
1662: of symmetry.
1663: 
1664: 
1665: 
1666: \end{appendix}
1667: 
1668:  
1669: 
1670: 
1671: \begin{thebibliography}{99}
1672: 
1673: \bibitem{Pie80} P.~Pieranski,
1674: % Two--dimensional interfacial colloidal crystals,
1675:          Phys.~Rev.~Lett. {\bf 45}, 569 (1980).
1676: 
1677: \bibitem{Ave02} %R.~Aveyard {\em et al.},
1678:  R.~Aveyard, B.~P.~Binks, J.~H.~Clint, P.~D.~I.~Fletcher, T.~S.~Horozov, B.~Neumann, V.~N.~Paunov, J.~Annesley, S.~W.~Botchway, D.~Nees, A.~W.~Parker, A.~D.~Ward, and A.~N.~Burgess,
1679: % Measurement of long-range repulsive forces between charged particles at an oil-water interface,
1680:           Phys.~Rev.~Lett. {\bf 88}, 246102 (2002).
1681: 
1682: \bibitem{Par07} B.~J.~Park, J.~P.~Pantina, E.~Furst, M.~Oettel, S.~Reynaert, and J.~Vermant,
1683: % Direct measurements of the effects of salt and surfactant on interaction forces between colloidal particles at water-oil interfaces,
1684:          Langmuir {\bf 24}, 1686 (2008).
1685: 
1686: \bibitem{Zen06} C.~Zeng, H.~Bissig, and A.~D.~Dinsmore,
1687: % Particles on droplets: From fundamental physics to novel materials,
1688:          Solid State Comm. {\bf 139}, 547 (2006).
1689: 
1690: \bibitem{Bre7r} F.~Bresme and M.~Oettel,
1691: % Nanoparticles at fluid interfaces,
1692:           J.~Phys.: Condens.~Matter {\bf 19}, 413101 (2007).
1693: 
1694: \bibitem{Oet08} M.~Oettel and S.~Dietrich,
1695: Langmuir {\bf 24}, 1425 (2008).
1696: 
1697: 
1698: \bibitem{Ghe97} F.~Ghezzi and J.~C.~Earnshaw,
1699: % Formation of meso--structures in colloidal monolayers,
1700:          J.~Phys.: Condens.~Matt. {\bf 9}, L517 (1997).
1701: 
1702: \bibitem{Gar98a} J.~Ruiz-Garc\'\i a and B.~I.~Ivlev,
1703: % Formation of colloidal clusters and chains at the air/water interface,
1704:          Mol.~Phys. {\bf 95}, 371 (1998).
1705: 
1706: \bibitem{Que01} %M.~Quesada--P\'erez {\em et al.},
1707:  M.~Quesada--P\'erez, A.~Moncho--Jord\'a, F.~Mart\'\i nez--L\'opez, and R.~Hidalgo--Alvarez,
1708: % Probing interaction forces in colloid monolayers: Inversion of structural data ,
1709:          J.~Chem.~Phys. {\bf 115}, 10897 (2001).
1710: 
1711: \bibitem{Gom05} O.~G\'omez--Guzm\'an and J.~Ruiz--Garc\'ia,
1712: % Attractive interactions between like-charged colloidal particles at the air/water interface,
1713:          J.~Colloid~Interface Sci. {\bf 291}, 1 (2005).
1714: 
1715: \bibitem{Che06} W.~Chen, S.~S.~Tan, Z.~S.~Huang, T.~K.~Ng, W.~T.~Ford, and P.~Tong,
1716: % Measured long--ranged attractive interaction between charged polystyrene latex spheres at a water-air interface,
1717:          Phys.~Rev.~E {\bf 74}, 021406 (2006).
1718: 
1719: \bibitem{Fer04} J.~C.~Fern\'andez--Toledano, A.~Moncho--Jord\'a, F.~Mart\'inez--L\'opez, and R.~Hidalgo--Alvarez,
1720: % Spontaneous formation of of mesostructures in colloidal monolayers trapped at the air-water interface: A simple explanation
1721:          Langmuir {\bf 20}, 6977 (2004).
1722: 
1723: \bibitem{Lav04}
1724:  I.~I.~Smalyukh, S.~Chernyshuk, B.~I.~Lev, A.~B.~Nych, U.~Ognysta, V.~G.~Nazarenko, and
1725:  O.~D.~Lavrentovich,
1726: % Ordered droplet structures at the liquid crystal surface and elastic-capillary colloidal interactions
1727:             Phys.~Rev.~Lett. {\bf 93}, 117801 (2004).
1728: 
1729: \bibitem{Lav07}
1730:  A.~B.~Nych, U.~M.~Ognysta, V.~M.~Pergamenshchik, B.~I.~Lev, V.~G.~Nazarenko, I.~Musevic, M.~Skarabot, and O.~D.~Lavrentovich,
1731: %  Coexistence of Two Colloidal Crystals at the Nematic-Liquid-Crystal–Air Interface,
1732:             Phys.~Rev.~Lett. {\bf 98}, 057801 (2007).
1733: 
1734: \bibitem{Kra00} P.~A.~Kralchevsky, K.~Nagayama,
1735: % Capillary interactions between particles bound to interfaces, liquid films and biomembranes,
1736:          Adv. Colloid Interface Sci. {\bf 85}, 145 (2000).
1737: 
1738: \bibitem{Nik02} %M.~G.~Nikolaides {\em et al.},
1739:  M.~G.~Nikolaides, A.~R.~Bausch, M.~F.~Hsu, A.~D.~Dinsmore, M.~P.~Brenner, C.~Gay, and D.~A.~Weitz,
1740: % Electric--field induced capillarity attraction between like--charged particles at liquid interfaces,
1741:          Nature {\bf 420}, 299 (2002).
1742: 
1743: \bibitem{Meg03} M.~Megens and J.~Aizenberg,
1744: % Like--charged particles at liquid interfaces: comment on Nikolaides et al., Nature {\bf 420} (2002) 299
1745:           Nature {\bf 424}, 1014 (2003).
1746: 
1747: \bibitem{Nik03} %M.~G.~Nikolaides {\em et al.},
1748:  M.~G.~Nikolaides, A.~R.~Bausch, M.~F.~Hsu, A.~D.~Dinsmore, M.~P.~Brenner, C.~Gay, and
1749: D.~A.~Weitz,
1750: % Electric--field induced capillarity attraction between like--charged particles at liquid interfaces: Reply on Comment,
1751:     Nature  {\bf 424}, 1014 (2003).
1752: 
1753: \bibitem{For04} L.~Foret and A.~W{\"u}rger,
1754: % Electric--field induced capillary interaction of charged particles at a polar interface,
1755:           Phys.~Rev.~Lett.~{\bf 92}, 058302 (2004). %cond-mat/0310657.
1756: 
1757: \bibitem{Kra04} K.~D.~Danov, P.~A.~Kralchevsky, and M.~P.~Boneva,
1758: % Electrodipping force acting on solid particles at a fluid interface,
1759:           Langmuir {\bf 20}, 6139 (2004).
1760: 
1761: \bibitem{Oet05} M.~Oettel, A.~Dom\'inguez, and S.~Dietrich,
1762: % Effective capillary interaction of spherical particles at  fluid interfaces,
1763:          Phys.~Rev.~E {\bf 71}, 051401 (2005). %cond-mat/0411329.
1764: 
1765: \bibitem{Oet05a} M.~Oettel, A.~Dom\'inguez, and S.~Dietrich,
1766: % Comment on Electrodipping force acting on solid particles at a fluid interface (Langmuir {\bf 20}, 6139 (2004).)
1767:          Langmuir {\bf 22}, 846 (2006).
1768: 
1769: \bibitem{Kra05}  K.~D.~Danov and P.~A.~Kralchevsky,
1770: % Reply to Comment
1771:          Langmuir {\bf 22}, 848 (2006).
1772: 
1773: \bibitem{Dan06} K.~D.~Danov, P.~A.~Kralchevsky, and M.~P.~Boneva,
1774: % Shape of the capillary meniscus around an electrically charged particle at a fluid interface: Comparison of theory and experiment,
1775:           Langmuir {\bf 22}, 2653 (2006).
1776: 
1777: \bibitem{Dom06} A.~Dom\'inguez,  M.~Oettel, and S.~Dietrich,
1778:          J.~Chem.~Phys. {\bf 128}, 114904 (2008).
1779: 
1780: \bibitem{Oet05b} M.~Oettel, A.~Dom\'inguez, and S.~Dietrich,
1781: % Attractions between charged colloids at water interfaces,
1782:          J.~Phys.: Condens.~Matter {\bf 17}, L337 (2005).
1783: 
1784: \bibitem{Wue05} A.~W{\"u}rger and L.~Foret,
1785: % Capillary atraction of colloidal particles at an aqueous interface,
1786:          J.~Phys.~Chem.~B {\bf 109}, 16435 (2005).
1787: 
1788: \bibitem{Dom05} A.~Dom\'inguez,  M.~Oettel, and S.~Dietrich,
1789: % Capillary induced interactions between colloids at an interface,
1790:          J.~Phys.: Condens.~Matter {\bf 17}, S3387 (2005).
1791: 
1792: \bibitem{Dom07} A.~Dom\'inguez,  M.~Oettel, and S.~Dietrich,
1793: % Theory of capillary-induced interactions beyond the superposition approximation
1794:          J.~Chem.~Phys. {\bf 127}, 204706 (2007).
1795: 
1796: \bibitem{deG74} P.~G.~de Gennes,
1797: {\em The Physics of Liquid Crystals} (Clarendon, Oxford 1993).
1798: 
1799: \bibitem{Lav03} M.~Kleman and O.~D.~Lavrentovich,
1800: {\em Soft Matter Physics: An Introduction} (Springer, New York 2003).
1801: 
1802: \bibitem{Sch98} J.~Schwinger, L.~L.~DeRaad Jr., K.~A.~Milton, and W.~Tsai,
1803: {\em Classical Electrodynamics} (Perseus, Reading 1998), Chap. 11.2.
1804: 
1805: \bibitem{Pou97} P.~Poulin, H.~Stark, T.~C.~Lubensky, and D.~A.~Weitz,
1806: % Novel colloidal interactions in anisotropic fluids,
1807:          Science {\bf 275}, 1770 (1997).
1808: 
1809: \bibitem{Lub98} T.~C.~Lubensky, D.~Pettey, N.~Currier, and H.~Stark,
1810: % Topological defects and interactions in nematic emulsions,
1811:          Phys.~Rev.~E {\bf 57}, 610 (1998).
1812: 
1813: \bibitem{Ruh97} R.~W.~Ruhwandl and E.~M.~Terentjev,
1814: % Long-range forces and aggregation of colloid particles in a nematic liquid crystal,
1815:          Phys.~Rev.~E {\bf 55}, 2958 (1997).
1816: 
1817: \bibitem{Ram96} S.~Ramaswamy, R.~Nityananda, V.~A.~Raghunathan, and J.~Prost,
1818: % Power-law forces between particles in a nematic,
1819:          Mol.~Cryst.~Liq.~Cryst. {\bf 288}, 175 (1996).
1820: 
1821: \bibitem{Fuk02} J.~Fukuda, B.~I.~Lev, K.~M.~Aoki, and H.~Yokoyama,
1822: % Interaction of particles in a deformed nematic liquid crystal,
1823:          Phys.~Rev.~E {\bf 66}, 051711 (2002).
1824: 
1825: \bibitem{Sta99} H.~Stark,
1826: % Director field configurations around a spherical particle in a nematic liquid crystal,
1827:          Eur.~Phys.~J.~B {\bf 10}, 311 (1999).
1828: 
1829: \bibitem{Tas05} D.~Andrienko, M.~Tasinkevych, and S.~Dietrich,
1830: % Effective pair interaction between colloidal particles at a nematic-isotropic interface, 
1831:          Europhys.~Lett. {\bf 70}, 95 (2005).
1832: 
1833: \bibitem{Pet98} D.~Pettey, T.~C.~Lubensky, and D.~R.~Link,
1834: % Topological inclusions in 2D semctic C films,
1835:          Liquid Crystals {\bf 25}, 579 (1998).
1836: 
1837: \bibitem{Lav07a} O.~Lavrentovich, private communication. 
1838: 
1839: \end{thebibliography}
1840: 
1841: 
1842: \end{document}
1843: