University of Ljubljana Faculty of Mathematics and Physics Department of Physics Boˇstjan Markun Casimir effect in smectic liquid crystals Doctoral thesis Adviser: prof. dr. Slobodan ˇ umer Ljubljana, 2007 Univerza v Ljubljani Fakulteta za matematiko in fiziko Oddelek za fiziko Boˇstjan Markun Casimirjev pojav v smektiˇcnih tekoˇcih kristalih Doktorska disertacija Mentor: prof. dr. Slobodan ˇ umer Ljubljana, 2007 Zahvaljujem se prof. Slobodanu ˇ umru za mentorsko vodstvo od samega zaˇcetka moje raziskovalne poti, katere rezultat je priˇcujoˇce delo. Sodelavcem Andreji, Brini, Danielu, Gregorju, Mateju in Mihu hvala za zelo prijetno delovno vzduˇsje in pomoˇc pri reˇsevanju takˇsnih in drugaˇcnih teˇzav. Se posebej sem hvaleˇzen Primoˇzu za vse pordeˇcele verzije mojih besedil ter za vse pametne in “pametne” nasvete. Disertacijo posveˇcam vsem svojim bliˇznjim, ki so mi vsa ta leta stali ob strani. Abstract The thesis deals with various aspects of the Casimir effect in smectic liquid crystals. The Casimir interaction in planar smectic-A systems is studied, considering both types of smectic ordering – positional and orientational – including the coupling between them. This provides a complete picture of the phenomenon in smectic-A systems with homogeneous equilibrium order. The behavior of the Casimir interac-tion in vicinity of the smectic-A to smectic-C phase transition is considered. The presence of this transition results in some special features of the interaction. A spe-cial attention is devoted to confined systems with non-trivial equilibrium order. The Casimir interaction in a homeotropic smectic cell with surface enhanced positional order is studied; an exponential decay of the Casimir force is predicted, contrary to the long-range interaction in homogeneous smectic systems. In a homeotropic nematic cell with surface induced presmectic order a faster decay of the Casimir force than in normal nematics is discovered. In addition, a few systems where inhomogeneity of equilibrium ordering does not affect the Casimir interaction are presented. Keywords: Casimir effect, smectic liquid crystals, fluctuations, confinement, phase transition PACS: 61.30.Dk, 61.30.Hn, 64.70.Md, 68.60.Dv Povzetek Delo je posveˇceno razliˇcnim vidikom Casimirjevega pojava v smektiˇcnih tekoˇcih kristalih. Obravnavamo Casimirjevo interakcijo v planarnih smektiˇcnih A sistemih. Pri tem zajamemo oba vidika smektiˇcne ureditve - pozicijskega in orientacijskega -ter sklopitev med njima. S tem podamo popoln opis Casimirjevega pojava v smektiˇcnih A sistemih s homogeno ravnovesno ureditvijo. Nadalje raziˇsˇcemo obnaˇsanje Casimirjeve sile v bliˇzini prehoda iz smektiˇcne A v smektiˇcno C fazo. Bliˇzina tega faznega prehoda se odraˇza v nekaterih posebnih lastnostih sile. Posebno pozornost v disertaciji posvetimo ograjenim sistemom z netrivialno ravnovesno ureditvijo. Obravnavamo homeotropno smektiˇcno celico s poveˇcanim povrˇsinskim pozicijskim redom. Ugotovimo, da je Casimirjeva sila v takem sistemu kratkega dosega, v nasprotju s homogenimi smektiˇcnimi sistemi, kjer je sila dolgega dosega. Izraˇcunali smo Casimirjevo silo v nematski homeotropni celici s povrˇsinsko vsiljenim pred-smektiˇcnim redom. V tem primeru sila upada precej hitreje kot v obiˇcajnem ne-matiku. Predstavili smo tudi nekaj sistemov, kjer nehomogena ravnovesna struktura ne vpliva na Casimirjevo interakcijo. Kljuˇcne besede: Casimirjev pojav, smektiˇcni tekoˇci kristali, fluktuacije, ograditev, fazni prehod PACS: 61.30.Dk, 61.30.Hn, 64.70.Md, 68.60.Dv Contents 1 Introduction 7 1.1 Casimir force in liquid crystals...................... 11 1.1.1 Nematic liquid crystals...................... 11 1.1.2 Smectic liquid crystals...................... 15 1.1.3 Search for experimental evidence ................ 18 1.1.4 Aim and outline of thesis..................... 20 2 Theoretical model 21 2.1 Free energy of smectic-A phase ..................... 21 2.2 Chiral smectics close to smectic-A* to smectic-C* phase transition . . 22 2.3 Confined systems............................. 26 2.4 Calculation of Casimir force....................... 27 3 Casimir force in smectic-A phase 33 3.1 Homeotropic smectic-A cell ....................... 33 3.1.1 Fluctuations of degree of smectic order ?............ 35 3.1.2 Fluctuations of director and smectic layers........... 36 3.2 Free-standing smectic-A film....................... 43 3.3 Casimir force in slightly dilated or compressed cell.......... 47 3.4 Importance of Casimir force in smectic-A systems........... 50 4 Casimir force in vicinity of smectic-A to smectic-C phase transition 53 4.1 Homeotropic cell............................. 54 4.1.1 Casimir force above Tc...................... 57 4.1.2 Casimir force in frustrated system (Tmax 0) the force should be calculated using the free energy F instead of the energy E as a proper thermodynamic potential. For high temperatures or large thicknesses (ksTh ^ hc), where the thermal fluctuations of the field Introduction 9 overwhelm the quantum fluctuations, the Casimir force reads F =------— =--------5~Cr(3) j (1.6) oh 47r/r where fcB is Boltzmann constant and (r is the Riemann zeta function with the value Cr(3) = 1.202.... It is worth mentioning that the electromagnetic Casimir force is a special case of the van der Waals force between two dielectric slabs interacting across another dielectric medium, first calculated by Lifshitz in 1955 [4]. As it can be inferred from the above derivation, the Casimir interaction is not specific only to the EM field. It is present in every confined system where the fluctuation spectrum of any physical field is modified due to some boundary conditions. In other words, it is omnipresent but not easily observed due to its usually small magnitude. We shall here give a brief overview of the various fields of physics where the Casimir force has been studied [2]. First of all, the original Casimir calculation, Eq. (1.5), has been refined and extended in various ways. As already mentioned, the contribution due to thermal fluctuations of electromagnetic field should be considered at finite temperatures, leading to the expression (1.6) in the limit of high temperatures [5-12]. Furthermore the corrections concerning finite conductivity and roughness of the plates have been evaluated [13-22]. The generalization to magnetically permeable plates has been performed, which can even change the sign of the force [23-27]. The Casimir interaction has been calculated for rectangular cavities and for spherical, cylindrical, toroidal and wedge geometries [28-34]. There seems to be no a-priori way to predict what the stress on specific geometrical object will be. For example, the Casimir force on the conducting spherical shell tends to expand it, contrary to the attraction obtained between two plates. Moreover, the interaction between the walls of a rectangular cavity can be either attractive or repulsive depending on the relationship between the lengths of the sides. The dynamical Casimir effect, describing the force and radiation from moving plates, has also received much attention [35-39]. In quantum field theory the Casimir effect has found application in the bag model of hadrons in quantum chromodynamics (QCD) [40-42] and in Kaluza-Klein field theories [40, 42-46]. Casimir-type effects naturally arise in cavity quantum electrodynamics (QED) [25] and even in electrical engineering of microchips [47-49]. In gravitational theory, cosmology and astrophysics the Casimir effect arises in space-times with nontrivial topology and is related to problems of particle creation by black holes, gravitational collapse and inflation process [42, 50-55]. As a mechanical analog, the acoustic Casimir force has to be mentioned. Larraza et al. managed to measure the force between two closely spaced plates due to the modification of the spectrum of acoustic noise [56-58]. The Casimir idea found application even in maritime physics, where an attractive force between two ships in a rough sea has been attributed to the modification of the wave spectrum in the region between the 10 Introduction ships [59]. A fluctuation-induced interaction is also present between inclusions in biological membranes. Here the thermal fluctuations of a membrane are hindered by the presence of the inclusions which leads to the interaction [60–62]. Many studies have been devoted to the thermal Casimir interaction in correlated fluids [63] – such as critical liquids and binary mixtures of liquids [64–69], super-fluids [70–73], liquid crystals and electrolytes [74–76]. The various studies of the Casimir effect in liquid crystals, the main topic of this thesis, are presented in detail later. Here we should mention a universal property of the Casimir force in planar ge-ometry which does not depend on details of the studied system: fluctuations with long-range correlations induce long-range interaction while short-correlated fluctua-tions result in a short-range force decaying with some characteristic length. Typical examples of long-range correlations include critical systems close to the phase tran-sition and systems with a massless Goldstone fluctuation modes due to the broken continuous symmetry of ordering. On the other hand, the sign of the Casimir force depends on the geometry and topology of the system as well as on the specific boundary conditions. The experimental studies of the Casimir force are vastly outnumbered by the theoretical work. The main reason lies in difficulty of experiments as the Casimir force is usually weak and often screened by other effects. The first documented suc-cessful attempts of measuring the electromagnetic Casimir force belong to Sparnaay in 1958 [77]. However, due to the poor accuracy of the measurements only quali-tative agreement with theoretical predictions was confirmed. A firm experimental measurement of the Casimir force was reported in 1997 by Lamoreaux [78, 79], almost half a century after the theoretical prediction. Lamoreaux used an elec-tromechanical system based on a torsion pendulum and measured the force between a gold-coated plate and sphere. The agreement with the theory was claimed to be within 5%. Subsequent experiments which relied on the atomic force microscopy (AFM) techniques [80–84] also produced results that were in excellent quantitative agreement with the theory. In recent years a number of new experiments ensued. The force between two crossed cylinders [85], plan-parallel plates [86] and in standard sphere-plane AFM setup was measured [87–91]. In dynamical experiments the influence of the Casimir force on the behavior of micromechanical oscillators was observed [92–95]. The precision of experiments has been greatly improved over the last years and now allows for delicate tests of the theoretical predictions. Further-more, the Casimir force measurements provide one of the most sensitive tests of the hypothetical new forces predicted by modern theories of fundamental interactions including corrections to Newtonian gravitational law at small distances [95–101]. Although the Casimir force is weak at macroscopic distances, it is important for modern technologies which involve ever smaller length scales where the Casimir force becomes dominant. It is presently unclear whether the Casimir force will present an obstacle or a useful feature in micro- and nano-engineering. For example, the Introduction 11 first microelectromechanical device which shows actuation by the Casimir force was designed by researchers at Bell Labs in 2001 [92]. On the other hand, the Casimir force restricts the yield and performance of nanoscale devices as the movable parts often stick together due to the strong attraction [102–104]. Apart from measurements of the electromagnetic Casimir force, the experiments on other systems have also been performed. The influence of the Casimir force has been observed in the wetting behavior of liquid helium on a metallic surface [70, 71]. The thickness of the helium film formed on the metal depends on the strength of the interaction between the surfaces of the film. When the system is cooled down to the fluid/super-fluid phase transition the fluctuations become critical and the magnitude of the Casimir force strongly increases which is reflected in thinning of the wetting film. Similar experiments were performed with binary liquid mixtures [67, 105–107] where a sharp increase of the wetting film thickness was observed near the critical (demixing) point due to the enhanced Casimir interaction. Casimir interaction is also expected to have an important role in physics of colloids where a long-range attraction would eventually lead to the flocculation of dispersed particles [69]. Such a flocculation of colloidal particles has actually been observed in binary liquid mixtures but the precise interpretation of experimental results is still unclear [108–110]. 1.1 Casimir force in liquid crystals Having briefly described the Casimir interaction in various fields of physics, ranging from biophysics to cosmology, we proceed with a thorough overview of the studies of the Casimir force in liquid-crystal systems. Liquid-crystalline phases are interme-diate states of matter between a liquid and a crystal phase [111]. They are formed by anisotropic molecules, usually elongated or disc-shaped. There exist a variety of distinct liquid-crystalline phases which are characterized by orientational and in some cases also by partial orientational order of constituent molecules. It has been established long ago in light-scattering experiments [112] that thermal fluctuations of ordering have an important role in liquid-crystal systems. These fluctuations are the source of the Casimir interaction when the system is confined by external boundaries. The richness of different phases, phase transitions, order parameters and couplings between them makes the liquid-crystalline systems especially attrac-tive for studying the phenomenology of the Casimir interaction. 1.1.1 Nematic liquid crystals Nematic phase is the simplest liquid-crystalline phase (Fig. 1.2). Molecules in a nematic phase are liquid-like in a sense that there is no long-range positional order and the translational motion of the molecules is random. However, there exists a long-range orientational order. Molecules tend to orient with their long axes parallel 12 Introduction n a -nv Figure 1.2 Nematic liquid crystal phase. The average orientation of molecules is described by a headless unit vector n called the director. The directions n and -n are physically equivalent. The angle ? gives the tilt of a molecule with respect to the director. The polar angle ? is used to describe biaxial ordering. to each other. This orientational order is described by director n, which is a unit vector giving the average local direction of orientation of molecules. The degree of orientational order is measured by the order parameter S = (3/2 cos2 ? — 1/2), where ? is the angle between the director n and long axes of the molecule, while the brackets denote the thermodynamic average. Nematic ordering is usually uniaxial, except in some special systems. The biaxial ordering is described by the biaxial director n^, perpendicular to n, and the degree of biaxiality P = (sin2(?) cos(2?)), where ? is the polar angle of molecular orientation. In equilibrium, with no external forces acting on the system, the director tends to be uniform over the whole sample. The energy cost of a director-field deformation is given by the Frank elastic free-energy [113] F = K1(V ¦ n)2 + K2(n • V x n)2 + K3(n 21 x V x n) J dV . (1.7) Here K1, K2 and K3 are splay, twist and bend elastic constants. A more complete description of nematic systems is given by the tensor order parameter Q which incorporates all aspects of nematic ordering - director (n), degree of ordering (S), and biaxiality (n&, P). The Q tensor is a traceless symmetric tensor based on some macroscopic quantity which is zero in the isotropic phase and non-zero in the nematic phase. The magnetic susceptibility tensor ? is usually used for this purpose and the order parameter tensor is defined as Q = C(? ~ 13I Tr?), where C is a normalization constant and I a unit tensor. The free energy density of a nematic system close to Introduction 13 the nematic-isotropic phase transition is then described by a Landau-type expansion f = —A(T - T*)TrQ2-----B TrQ3 + -C(TrQ2)2 +L?Q.?Q , (1.8) where A, B, C and L are material constants and T* is supercooling limit of isotropic phase. The first study of the Casimir force in a nematic system was performed by Ajdari et al. in 1991 [114, 115]. They calculated the force in a nematic homeotropic cell (Fig. 1.3), consisting of two infinite parallel plates separated by the distance h which enforce homeotropic orientation of the director [n(z = 0) = n(z = h) = (0,0,1)]. These imposed boundary conditions hinder the thermal fluctuations of the nematic 1 I Figure 1.3 Homeotropic nematic cell with the director structure n = nz. The arrows indicate the enforced orientation of the director at the plates. director in the cell, thus modifying the spectrum of fluctuations which leads to the Casimir interaction. The Casimir force in this configuration is equal to kBTS (K3 K3 \ FCas =-------t^"?r(3) -K~ + tt- . (1.9) 8?h3 1 K2 We note here that this force is equal to the thermal EM Casimir force between two metallic plates [Eq. (1.6)], apart from factor including the ratio of elastic constants. This demonstrates the universality of the Casimir interaction which does not depend on specific details of the studied system but on the type of fluctuation modes and on imposed boundary conditions. We also note that director fluctuations in nematic liquid crystals are an example of massless Goldstone fluctuation modes, which try to recover broken continuous symmetry of a high temperature - in this case isotropic - phase. The work of Ajdari was extended by Ziherl et al. [116] who evaluated the contributions of fluctuations of biaxiality and degree of nematic order to the Casimir force. These contributions are equal to kTS 1 ^^ exp (-2hk/?i) /1 h h 2 2 X ( 1 h h 2\ 2 ?i ?i FCas =----:—tt / --t-------- +----k + 2 k , (1.10) 4? h3 -^—' k3 2 ?i rjf where ?i are the corresponding correlation lengths of fluctuations. In the limit of large thicknesses (h/?i ^ 1) this force decays as exp(-2h/?i)/h. This is a 14 Introduction demonstration of another universal feature of the Casimir force - massive short-range correlated fluctuations result in a short-range Casimir force, decaying exponentially with some characteristic length equal to the correlation length of fluctuation modes in question. It was established that these short-range contributions are important only close to the nematic-isotropic phase transition where the correlation lengths of the massive modes are strongly increased. Otherwise the long-range contribution of the director fluctuations [Eq. (1.9)] dominates the Casimir force in nematics. These basic results [Eqs. (1.9,1.10)] were also generalized for finite anchoring strengths, where the ordering at the plates is not fixed but can deviate from a preferred value [116, 117]. It was established that finite surface coupling reduces the magnitude of the Casimir force and can even modify its thickness dependence and sign. It is interesting to note that in the case of no surface coupling between the plates and liquid crystal the Casimir force is exactly the same as in the case of infinitely strong anchoring. These two limiting cases correspond to the so-called Neumann and Dirichlet boundary conditions, respectively. The structure of the eigen-modes in the two cases is different, but the energy spectra are identical which leads to identical Casimir forces. Even more interesting is the case of mixed (Dirichlet-Neumann) boundary conditions where the force changes the sign and becomes repulsive [114, 117]. Further studies addressed different aspects of the Casimir force in nematic liquid crystals. Li and Kardar evaluated corrections to the Casimir force due to the roughness of the plates [118, 119]. Ziherl et al. studied the force in a pre-nematic wetting system with inhomogeneous equilibrium order [120]. They found that the Casimir force in such a system is repulsive and short-range. Much attention has been paid to the so-called frustrated systems such as the hybrid and Freedericksz cell [121]. The hybrid cell is similar to the homeotropic cell (Fig. 1.3) except that now one I h 0. The rest gives the various contributions to elastic free energy weighted by the elastic constants C\\, d1, d2 and d3. The subscripts || and _L denote directions parallel and perpendicular to the layer normal, respectively. We shall seldom employ the elastic contributions in full generality as the lowest order elastic terms are usually sufficient to describe smectic systems. The second part fN to the total smectic free energy comes from the energy cost of deformations of orientational order and is given by the usual nematic Frank elastic energy: fN = -K1(V ¦ n) 2 + K 2[n • (V x n)] 2 +K3[n x (V x n)] 2 . (2.2) 2 2 2 Here K1, K2 and K3 are splay, twist and bend elastic constants, respectively. The third contribution fLN describes the coupling between the orientational and the positional order. In the case of the smectic-A phase it is given by 1 2 fln = ~C± (Vj_ + iq0?n±)? . (2.3) 2 21 22 Theoretical model In the case of the smectic-C phase a fourth order term should be included so that 1 (1) 2 1 (2) fln = —C]_ (Vj_ + iqo?n±)? + —C]_ (Vj_ + iqo?n±)? , (2.4) 4 with C\_ < 0 and C\_ > 0, but we refer to this case in a more specific system later on. The complete free energy can now be written as F = (fl + fn + fln)dV. (2.5) It is useful to expand the free energy density in terms of ? and ?. We obtai ain 11 1 fl = -a? + -b? +C\\ (V||? ) + 2 42 1 2 2 2n C|| |(V||? ) +? (V||?) J +d\\ \V2 ? — ? (V ±? ) 2] + 2X?±? ¦ J\7±? + ? (V2±? ) 2 1 + d 2 l^2ip — ? (V\\? ) 2] + \2X7\\?J\7\\? + ?V2 ?] 2 I L " N II II II 1 f L II II ? L II II II ) in 2 ? + [ 2 -^ 2 (2.6) 1 fln = ~C± (v±?) + ? Vi? + qo?n_i_ . (2.7) 2 v If no elastic deformations are present in the system, then the equilibrium value of bulk smectic order is given by ?o = \J—a/b, the phase ? is constant [? = ?(r)] and the director is perpendicular to the layers (?n = 0). Above the phase transition temperature TNa in nematic phase there is no bulk smectic order and ? = 0. 2.2 Chiral smectics close to smectic-A* to smectic-C* phase transition In this thesis, we also address the behavior of the Casimir force close to the smectic-A to smectic-C phase transition. We actually also consider the more complex chiral smectic phases, smectic-A* (Sm-A*) and smectic-C* (Sm-C*), as this brings no conceptual difficulties to our calculations and the results can be straightforwardly applied to the non-chiral phases. Chiral molecules, which lack the mirror symmetry and therefore distinguish left-and right-handed types, form chiral liquid crystal phases [158-160]. The chiral Sm-A* phase exhibits the same structure as its non-chiral counterpart, but its physical properties are different. On the other hand, the chirality modifies the structure of Sm-C* phase. The molecules in Sm-C* are still tilted with respect to the layer normal as in a non-chiral Sm-C phase. However the direction of the tilt changes gradually from layer to layer such that the director forms a helical structure (Fig. 2.1). Theoretical model 23 Figure 2.1 Helical structure of chiral Sm-C? phase (the period of the he-lix on the figure is exaggeratedly short). The direction of the molecular tilt changes gradually from layer to layer. The arrows indicate the orientation of spontaneous polarization. The period of the helix (? 1 µm) is incommensurate with the layer thickness and much larger compared to it. Furthermore, the Sm-C* phase can posses spontaneous polarization and is hence ferroelectric. This was discovered experimentally by Meyer et al. in 1975 [161] and later explained by Meyer on pure symmetry grounds [162]. The spontaneous polarization is oriented perpendicular both to the layer normal and to the director. The ordering in Sm-A* and Sm-C* phases can be described by two two-component order parameters. The primary order parameter, L = (?x,?y), represents the projection of the director onto the x-y plane. The secondary order parameter is the spontaneous polarization P = (Px, Py). However, the interaction between the dipole moments of molecules is too weak to drive the Sm-A* - Sm-C* phase transition. In Sm-C* phase, the polarization P appears only due to the coupling with the tilt L and for this reason smectics belong to the class of so-called improper ferroelectrics. The Sm-A* - Sm-C* phase transition can be conveniently described by a phe-nomenological Landau-type model. In this model, the free energy density reads 1 f = fa+a 2 ?x 2 + ? 2y 1 +K\ 2 1 2? + 1 + —b 4 1 ?x dy +K2 2 2A2 ? ??y ?x + ?y ) - ? 4a; -7— \ ?z . 2 / ??x ?? y\ 1 K2 -7:----------7T~ + - ?y ?x 2 2 ?? ?z ?y ?x ?z ?x—------?y 1 +Ks ??x ?z ??x x2 , ,21 ??x \ ( ??y \ —— + -7— ?z ?z (2.8) 1 , 2 2A ??x ??y +-----( Px + Py ) - µ Px ?------ Py^Z— + C (Px ?y - Py ?x) . 2? z ?z Here f a stands for the equilibrium free energy density of Sm-A* phase. The temper- 24 Theoretical model Figure 2.2 Order parameters of the Sm-A* - Sm-C* phase transition: a) tilt of molecules L ; b) spontaneous polarization P. Molecular orientation can be also described by the azimuthal angle 9 and the polar angle