Skip to main content

Hydrodynamics and Electrohydrodynamics of Liquid Crystals

  • Chapter

Part of the book series: Partially Ordered Systems ((PARTIAL.ORDERED))

Abstract

We present the hydrodynamic and electrohydrodynamic equations for uniaxial nematic liquid crystals and explain their derivation in detail. To derive hydrodynamic equations, which are valid for sufficiently small frequencies in the limit of long wavelengths, one identifies first the hydrodynamic variables, which come in two groups: quantities obeying conservation laws and variables associated with spontaneously broken continuous symmetries. As variables that characterize the spontaneously broken continuous rotational symmetries of a nematic liquid crystal we have the deviations from the preferred direction, which is characterized by the director, a unit vector that does not distinguish between head and tail.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   84.99
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Preview

Unable to display preview. Download preview PDF.

Unable to display preview. Download preview PDF.

References

  1. L.P. Kadanoff and P.C. Martin, Ann. Phys. 24, 419 (1963).

    MathSciNet  ADS  MATH  Google Scholar 

  2. P. Hohenberg and P.C. Martin, Ann. Phys. 34, 291 (1965).

    ADS  Google Scholar 

  3. LM. Khalatnikov, Introduction to the Theory of Superfluidity, Benjamin, New York (1965).

    Google Scholar 

  4. N.N. Bogoljubov, Phys. AbhandL SU 6, 229 (1962).

    Google Scholar 

  5. P.C. Martin, O. Parodi, and P. Pershan, Phys. Rev. A6, 2401 (1972).

    ADS  Google Scholar 

  6. D. Forster, Hydrodynamic Fluctuations, Broken Symmetry and Correlation Functions, Benjamin, Reading, MA. (1975).

    Google Scholar 

  7. T.C. Lubensky, Phys. Rev. A6, 452 (1972).

    ADS  Google Scholar 

  8. M.J. Stephen and J.P Straley, Rev. Mod. Phys. 46, 617 (1974).

    ADS  Google Scholar 

  9. H.R. Brand and H. Pleiner, J. Phys. (Paris) 41, 553 (1980). Unfortunately, this paper contains various misprints and errors.

    MathSciNet  Google Scholar 

  10. H. Pleiner and H.R. Brand, Phys. Rev. A25, 995 (1982).

    ADS  Google Scholar 

  11. H.R. Brand and H. Pleiner, Phys. Rev. A24, 1783 (1981);

    Google Scholar 

  12. H.R. Brand and H. Pleiner, Phys. Rev. A26, 1783 (1982),

    ADS  Google Scholar 

  13. H.R. Brand and H. Pleiner, Phys. Rev. A30, 1548 (1984).

    Google Scholar 

  14. W.M. Saslow, Phys. Rev. A25, 3350 (1982).

    MathSciNet  ADS  Google Scholar 

  15. M. Liu, Phys. Rev. A24, 2720 (1981).

    ADS  Google Scholar 

  16. H.R. Brand and H. Pleiner, J. Phys. (Paris) 45, 563 (1984).

    Google Scholar 

  17. H. Pleiner and H.R. Brand, Phys. Rev. A29, 911 (1984) and H. Pleiner, Mol. Cry st. Liq. Cryst. 114, 103 (1984).

    ADS  Google Scholar 

  18. H. Pleiner and H.R. Brand, Phys. Rev. A29, 911 (1984) and H. Pleiner, Mol. Cry st. Liq. Cryst. 114, 103 (1984).

    Google Scholar 

  19. H.R. Brand and H. Pleiner, J. Phys. (Paris) II 1, 1455 (1991).

    Google Scholar 

  20. M. Liu, Phys. Rev. A19, 2090 (1979).

    ADS  Google Scholar 

  21. K.A. Hossain, J. Swift, J.-H. Chen, and T.C. Lubensky, Phys. Rev. B19, 432 (1979).

    ADS  Google Scholar 

  22. B.S. Andereck and J. Swift, Phys. Rev. A25, 1084.

    Google Scholar 

  23. H.R. Brand and H. Pleiner, J. Phys. (Paris) 43, 853 (1982).

    MathSciNet  Google Scholar 

  24. H.R. Brand and P. Bak, Phys. Rev. A27, 1062 (1983).

    ADS  Google Scholar 

  25. H.R. Brand and J. Swift, J. Phys. Lett. (Paris) 44, 333 (1983).

    ADS  Google Scholar 

  26. H.R. Brand, Mol Cryst. Liq. Cryst. Lett. 4, 23 (1986).

    Google Scholar 

  27. H.R. Brand, Phys. Rev. A33, 643 (1986).

    ADS  Google Scholar 

  28. H.R. Brand and H. Pleiner, J. Phys. Lett. (Paris) 46, L 711 (1985).

    Google Scholar 

  29. H. Pleiner and H.R. Brand, Phys. Rev. A39, 1563 (1989).

    ADS  Google Scholar 

  30. H.R. Brand and K. Kawasaki, J. Phys. (Paris) II 2, 1789 (1992).

    Google Scholar 

  31. M. Kléman, Rep. Progr. Phys. 52, 1455 (1989).

    Google Scholar 

  32. H. Pleiner and H.R. Brand, Phys. Rev. Lett. 54, 1817 (1985).

    ADS  Google Scholar 

  33. H. Pleiner, Liq. Cryst. 1, 197 (1986).

    Google Scholar 

  34. H. Pleiner, Liq. Cryst. 3, 249 (1987).

    Google Scholar 

  35. H. Pleiner. Phys. Rev. A37, 3986 (1988).

    ADS  Google Scholar 

  36. H.R. Brand, P.E. Cladis, and H. Pleiner, Macromol. 25, 7223 (1992).

    ADS  Google Scholar 

  37. H. Pleiner and H.R. Brand, Mol Cryst. Liq. Cryst. 199, 407 (1991).

    Google Scholar 

  38. H. Pleiner and H.R. Brand, Macromol. 25, 895 (1992).

    ADS  Google Scholar 

  39. H.R. Brand and H. Pleiner, Physica A 208, 359 (1994).

    ADS  Google Scholar 

  40. C.W. Oseen, Trans. Far. Soc. 29, 883 (1933).

    Google Scholar 

  41. H. Zöcher, Trans. Far. Soc. 29, 945 (1933).

    Google Scholar 

  42. F.C. Frank, Disc. Far. Soc. 25, 19 (1958).

    Google Scholar 

  43. H. Pleiner and H.R. Brand, J. Phys. Lett. (Paris) 41, L 491 (1980).

    MathSciNet  Google Scholar 

  44. W. Helfrich, J. Chem. Phys. 51, 4092 (1969).

    ADS  Google Scholar 

  45. E. Dubois-Violette, G. Durand, E. Guyon, P. Manneville and P. Pieranski, in Solid State Physics, Suppl. 14, ed. by L. Liébert, Academic, New York (1978).

    Google Scholar 

  46. E. Bodenschatz, W. Zimmermann, and L. Kramer, J. Phys. (Paris) 49, 1875 (1988).

    Google Scholar 

  47. R.B. Meyer, Phys. Rev. Lett. 22, 918 (1969).

    ADS  Google Scholar 

  48. P.G. de Gennes and J. Prost, The Physics of Liquid Crystals, 2nd ed., Clarendon, Oxford University Press (1993).

    Google Scholar 

  49. H.R. Brand and H. Pleiner. Phys. Rev. A35, 3122 (1987).

    ADS  Google Scholar 

  50. H. Pleiner and H.R. Brand, Phys. Rev. A36, 4056 (1987).

    ADS  Google Scholar 

  51. H.R. Brand and H. Pleiner, Phys. Rev. A37, 2736 (1988).

    ADS  Google Scholar 

  52. H.R. Brand and H. Pleiner, Mol. Cryst. Liq. Cryst. Lett. 8, 11 (1991).

    Google Scholar 

  53. H. Pleiner and H.R. Brand, Phys. Rev. A43, 7064 (1991).

    ADS  Google Scholar 

  54. H. Pleiner and H.R. Brand, Mol. Cryst. Liq. Cryst. 257, 289 (1994).

    Google Scholar 

  55. H.R. Brand and H. Pleiner, Mol Cryst. Liq. Cryst. Lett. 226, 189 (1993).

    Google Scholar 

  56. H.R. Brand and H. Pleiner, J. Phys. (Paris) II 2, 1909 (1992).

    Google Scholar 

  57. J.L. Ericksen, Arch. Rat. Mech. Anal. 4, 231 (1960);

    MathSciNet  MATH  Google Scholar 

  58. J.L. Ericksen, Arch. Rat. Mech. Anal. 19, 357 (1966).

    Google Scholar 

  59. F.M. Leslie, Arch. Rat. Mech. Anal. 28, 265 (1968).

    MathSciNet  MATH  Google Scholar 

  60. J. Jähnig and H. Schmidt, Ann. Phys. 71, 129 (1972).

    ADS  Google Scholar 

  61. J.D. Lee and A.C. Eringen, J. Chem. Phys. 54, 5027 (1971).

    ADS  Google Scholar 

  62. Summation over repeated (Cartesian) indices is always assumed. Some formulas are written using (Cartesian) indices for clarity; they can always be rewritten in a coordinate free form by returning to the vector — tensor notation.

    Google Scholar 

  63. Here ∈ ijk is the totally antisymmetric third rank tensor.

    Google Scholar 

  64. Since the stress tensor is not completely defined by Eq. (2.4), one can make use of this freedom in order to redefine any stress tensor that fulfills condition (2.11) by \( {\tilde{\sigma }_{{ij}}} \equiv {\sigma_{{ij}}} + {\nabla_l}({\phi_{{jli}}} + {\phi_{{ilj}}} - {\phi_{{ijl}}}) \) resulting in a symmetric stress tensor \( ({\tilde{\sigma }_{{ij}}} = {\tilde{\sigma }_{{ji}}}) \) and leaving Eq. (2.4) unchanged \( ({\nabla_j}{\sigma_{{ij}}} = {\nabla_j}{\tilde{\sigma }_{{ij}}}) \).

    Google Scholar 

  65. H.R. Brand and H. Pleiner, J. Phys. Lett. (Paris) 42, L 327 (1981).

    Google Scholar 

  66. H. Pleiner and H.R. Brand, J. Phys. (Paris) 46, 615 (1985).

    Google Scholar 

  67. In a biaxial nematic system with two independent preferred directions, rotational symmetry would be broken completely.

    Google Scholar 

  68. If the broken symmetry was discrete, the different degenerate ground states would be separated by energy walls, making zero frequency switching between different ground states impossible.

    Google Scholar 

  69. Usually the temporal changes of symmetry variables show a simple behavior under the symmetry they are related to. Very often this can be described using the Poisson bracket formalism. Generally this method, however, is insufficient to give the full (reversible) dynamics of the symmetry variable. This is different in Mori’s projector formalism as applied to hydrodynamics by D. Forster [6]. In this case it becomes clear that for the linearized hydrodynamic equations both, the instantaneous and the noninstantaneous response, contribute to reversible hydrodynamics.

    Google Scholar 

  70. H. Pleiner in Incommensurate Crystals, Liquid Crystals, and Quasicrystals, edited by J.F. Scott and N.A. Clark, Plenum, New York (1987), p. 241.

    Google Scholar 

  71. H.R. Brand, H. Pleiner and W. Renz, J. Phys. (Paris) 51, 1065 (1990).

    Google Scholar 

  72. H. Pleiner and H.R. Brand, Europhys. Lett. 15, 393 (1991).

    ADS  Google Scholar 

  73. In order to get bulk quantities the (volume) densities have to be multiplied by the volume V, while concentration and director rotations must be multiplied by the total mass (= ρV).

    Google Scholar 

  74. If c describes the concentration of component 1 (c ≡ ρ 1 with ρ = ρ 1 + ρ 2 ), then the relative chemical potential is related to the individual chemical potentials by \( {\tilde{\mu }_c} = \rho {\mu_c} = \rho ({\mu_1} - {\mu_2}) \) and for the “chemical potential” µ one finds ρµ = µ 1 ρ 1 + µ 2 ρ 2 +h i δn i .

    Google Scholar 

  75. We note that by (2.20) h i , is defined differently from Ref. 45 by a global minus sign in accordance with Ref. 6. The variational derivative with respect to one variable is taken, while all other variables are kept at a fixed value. This will always be the case throughout the rest of this chapter.

    Google Scholar 

  76. H.B. Callen, Thermodynamics, 1st ed., John Wiley, New York (1960) and 2nd ed. (1985).

    MATH  Google Scholar 

  77. There exist additional surface contributions [usually called K 24 and K 13 terms, compare to J. Nehring and A. Saupe, J. Chem. Phys. 54, 337 (1971)], which are proportional to Ŝ i n i j n j or Ŝn j or Ŝ i n j j n i (Ŝ is the surface normal). To deal with realistic surface problems, however, very often one has to take into account surface orienting energies additionally, which involve the director orientation only (not its gradients) depending generally on (n - n s )2, where n s is the preferred director orientation at the surface. We will not consider surface effects here.

    ADS  Google Scholar 

  78. H. Pleiner and H.R. Brand, Europhys. Lett. 9, 243 (1989).

    ADS  Google Scholar 

  79. Some of these conditions are redundant, since with a and c - b 2 /a positive, c is also positive for real coefficients a, b, and c. This remark also applies to positivity relations given below (for static susceptibilities as well as for dynamic transport parameters).

    Google Scholar 

  80. The mass current g = ρ v has no irreversible part, since the total mass can only be transported by flow.

    Google Scholar 

  81. This justifies a posteriori the choice for the convective term made in the dynamical equation for the director (2.15) (Ref. 9).

    Google Scholar 

  82. In Eqs. (2.35)–(2.37) the orthogonality constraint (2.14) is explicitly built in by using the transverse Kronecker delta.

    Google Scholar 

  83. In the Leslie-Ericksen description, where the distinction between reversible and irreversible dynamics is not made, the parameter λ is obtained as a ratio of two dissipative transport parameters.

    Google Scholar 

  84. Since there is no irreversible part of g, the gradient of the chemical potential \( \vec{\nabla }\mu \) cannot act as a thermodynamic force of an irreversible process.

    Google Scholar 

  85. S.R. deGroot and P. Mazur, Nonequilibrium Thermodynamics, 2nd ed., Dover, New York (1984).

    Google Scholar 

  86. One should not mix up the dielectric tensor ∈ ij and its eigenvalues ∈, ∈ or ∈a with the energy density , nor the electric field vector E or its components E i with the (total) energy E. We are using Gaussian units. The conversion rules to MKSA units are listed at the end of the Appendix.

    Google Scholar 

  87. There is also no longer a true phase transition between the “isotropic” phase (no director) and the nematic phase (where the director exists), since the external field already defines a preferred direction in both phases. The difference is only quantitative, where in the “isotropic” phase the strength of the orientational order is very small but strong in the nematic one. This is quite analogous to the paramagnetic to ferromagnetic “phase transition” in the presence of an external magnetic field.

    Google Scholar 

  88. A similar situation occurs in superfluid 3He—A, where a small symmetry breaking energy contribution already exists intrinsically, compare to R. Graham and H. Pleiner, Phys. Rev. Lett. 34, 792 (1975).

    ADS  Google Scholar 

  89. If one wants to extend the dynamic description to very high frequencies, of course the full Maxwell equations have to be used including curl H and curl E (or rather the vector potential A) as nonhydrodynamic variables, compare to M. Liu, Phys. Rev. Lett. 70, 3580 (1993) and H.R. Brand and H. Pleiner, Phys. Rev. Lett. 74, 1883 (1995).

    ADS  Google Scholar 

  90. If one wants to extend the dynamic description to very high frequencies, of course the full Maxwell equations have to be used including curl H and curl E (or rather the vector potential A) as nonhydrodynamic variables, compare to M. Liu, Phys. Rev. Lett. 70, 3580 (1993) and H.R. Brand and H. Pleiner, Phys. Rev. Lett. 74, 1883 (1995).

    ADS  Google Scholar 

  91. Sometimes it is more suitable to write \( E = {E_0} - \vec{\nabla }\Phi ' \) where E 0 is the external field and Ф’ is the potential due to internal charges only.

    Google Scholar 

  92. The isotropic part of the Maxwell stress tensor is already part of the definition (2.52) of the pressure.

    Google Scholar 

  93. K. Henjes and M. Liu, Ann. Phys. 223, 243 (1993) and

    MathSciNet  ADS  MATH  Google Scholar 

  94. K. Henjes, Ann. Phys. 223, 277(1993).

    MathSciNet  ADS  MATH  Google Scholar 

  95. In setting up the Gibbs relation (2.51) and the pressure (2.52) as well as the momentum balance (2.59) we have neglected all contributions of order v;/c [compare to S.R. de Groot and L.G. Suttorp, Foundations of Electrodynamics, North-Holland, Amsterdam (1972)], since the typical hydrodynamic velocities v; are much smaller than the velocity of light c. The same consideration allows us to keep the electric degree of freedom while discarding simultaneously the magnetic degree of freedom.

    Google Scholar 

  96. We have made use of the quasistatic condition (2.49) in order to reduce the number of transport parameters from two to one.

    Google Scholar 

  97. Inserting the charge conservation law (2.12) and the entropy balance (2.40) into the Gibbs relation (2.51) the electric part of the dissipation function is R(E) = E . j’.

    Google Scholar 

  98. C.W. Gardiner, Handbook of Stochastic Methods, Springer Berlin, 2nd ed. (1985).

    Google Scholar 

  99. L.D. Landau and E.M. Lifshitz, Statistical Physics, Pergamon New York (1978).

    Google Scholar 

  100. L.D. Landau and E.M. Lifshitz, Fluid Mechanics, Pergamon Press, New York (1982).

    Google Scholar 

  101. Because of the definition (2.68) for the generalized forces, each dissipative transport parameter carries an explicit factor \( \gamma_{{\alpha \beta }}^{{diss}} = T\tilde{\gamma }_{{\alpha \beta }}^{{diss}} \) and the fluctuation dissipation theorem can be written in the form \( c\alpha \beta = {k_B}T\tilde{\gamma }_{{\alpha \beta }}^{{diss}} \).

    Google Scholar 

  102. In the case of (2.70i), higher order gradient contributions to the dissipation function have to be considered (cf. Section 2.6.5); \( {\vec{\nabla }^{{(1)}}} \) means ∂/r 1.

    Google Scholar 

  103. H. Pleiner and H.R. Brand, Phys. Rev. A27, 1177 (1983).

    ADS  Google Scholar 

  104. H. Pleiner and H.R. Brand, J. Phys. Lett. (Paris) 44, L 23 (1983).

    Google Scholar 

  105. P.C. Martin, Measurements and Correlation Functions, Gordon and Breach, New York (1968).

    Google Scholar 

  106. Simply adding fluctuating terms to nonlinear differential equations is no valid procedure [compare to N.G. van Kampen, Stochastic Processes in Physics and Chemistry, North-Holland, Amsterdam (1981)] and to date no physically reasonable and mathematically sound set of nonlinear fluctuating hydrodynamic equations is known, even for simple liquids.

    MATH  Google Scholar 

  107. R. Alben, Mol. Cryst. Liq. Cryst. 10, 21 (1970).

    Google Scholar 

  108. M.J. Freiser, Phys. Rev. Lett. 24, 1041 (1970).

    ADS  Google Scholar 

  109. A. Wulf, J. Chem. Phys. 59, 6596 (1973).

    ADS  Google Scholar 

  110. R.G. Priest and T.C. Lubensky, Phys. Rev. A9, 893 (1974).

    ADS  Google Scholar 

  111. J.P. Straley, Phys. Rev. A10, 1881 (1974).

    ADS  Google Scholar 

  112. Even if the second preferred direction m’ is not orthogonal to n, one can always construct a tripod of three mutually orthogonal unit vectors n, mp × n and p with p ≡ (n × m’)/ ∣n × m’∣.

    Google Scholar 

  113. For a partial lifting of this constraint in a mixture of two uniaxial nematics with different preferred directions cf. ref. [62].

    Google Scholar 

  114. The highest possible biaxial symmetry is orthorhombic (three mutual orthogonal twofold rotational symmetry axes). If the two preferred directions are not orthogonal, a lower symmetry is obtained. If there are more than two preferred directions (still breaking rotational symmetry completely), a symmetry higher than orthorhombic and even a noncrystallographic symmetry can result, compare to M. Liu, Phys. Rev. A24, 2720 (1981).

    ADS  Google Scholar 

  115. If a partial integration of ∫ ∈(flexo) dV is done, exactly the same expression as in (2.73) is obtained. Thus, all six flexoelectric coefficients are bulk parameters and their number is not reduced by the electrostatic condition (2.50), since ∇ j E i never occurs in Eq. (2.73). If linearized around the true equilibrium state (but not around a stationary nonequilibrium one), only the the symmetric part \( {\tilde{e}_{{i\alpha k}}} + {\tilde{e}_{{k\alpha i}}} \) contributes. This phenomenon is quite similar to the uniaxial nematic case, where generally \( {\tilde{e}_1} \ne {\tilde{e}_3} \) [35], although in a linearized equilibrium theory only \( {\tilde{e}_1} + {\tilde{e}_3} \) occurs.

    Google Scholar 

  116. In (2.75) the electrostatic condition (2.50) directly applies and reduces the number of independent parameters.

    Google Scholar 

  117. N.D. Mermin and T.-L. Ho, Phys. Rev. Lett. 36, 594 (1976).

    ADS  Google Scholar 

  118. A. Saupe, J. Chem. Phys. 75, 5118 (1981).

    ADS  Google Scholar 

  119. L.J. Yu and A. Saupe, Phys. Rev. Lett. 45, 1000 (1980).

    ADS  Google Scholar 

  120. F. Hessel and H. Finkelmann, Polym. Bull. (Berlin) 14, 375 (1985).

    Google Scholar 

  121. S.M. Fan, I.D. Fletcher, B. Gründogan, N.J. Heaton, G. Kothe, G.R. Luckhurst, and K. Praefcke, Chem. Phys. Lett. 204, 517 (1993).

    ADS  Google Scholar 

  122. P.G. de Gennes, Solid State Commun. 10, 753 (1972).

    ADS  Google Scholar 

  123. S. Dumrograttana, C.C. Huang, G. Nounesis, S.C. Lien, and J.M. Viner, Phys. Rev. A34, 5010 (1986);

    ADS  Google Scholar 

  124. P.M. Chaikin and T.C. Lubensky, Principles of Condensed Matter Physics, Cambridge University, Boston (1992).

    Google Scholar 

  125. They are used as well for describing defects and their dynamics, [compare to [30] and H. Pleiner, Phil. Mag. A 54, 421 (1986)].

    ADS  Google Scholar 

  126. Since in the isotropic phase the equilibrium value of S is zero, δS is identical to S itself.

    Google Scholar 

  127. H.R. Brand, Mol. Cry st. Liq. Cryst. Lett. 3, 147 (1986).

    MathSciNet  Google Scholar 

  128. H.R. Brand and K. Kawasaki, J. Phys. C 19, 937 (1986).

    ADS  Google Scholar 

  129. The dynamics of any microscopic scalar variable would be isomorphic to that of δS. There is only a quantitative difference in the relaxation times, which justifies keeping δS as variable under certain conditions, while discarding all other microscopic degrees of freedom.

    Google Scholar 

  130. In deriving Eq. (2.85) c2- c 20 ≪ c 20 has been assumed. For ωτ ≪ 1 the angle dependent part of (2.85) becomes imaginary and contributes to D(∅) only.

    Google Scholar 

  131. M.E. Mullen, B. Lüthi and M.J. Stephen, Phys. Rev. Lett. 28, 799 (1972).

    ADS  Google Scholar 

  132. In Ref. 34 special cases of Eqs. (2.85) and (2.86) are given. Using a scalar order parameter for a macroscopic dynamics near the nematic to smectic A phase transition, rather similar features in the sound mode spectrum are found, compare to M. Liu, Phys. Rev. A19, 2090 (1979).

    ADS  Google Scholar 

  133. H. Finkelmann, H. Ringsdorf, and J.H. Wendorff, Makromol. Chem. 179, 273 (1978).

    Google Scholar 

  134. In this respect we have changed the point of view expressed in Ref. 67.

    Google Scholar 

  135. J.K. Krüger, C. Grammes, and J.H. Wendorff, Progr. Colloid and Polymer Sci. 80, 45 (1989).

    Google Scholar 

  136. F.W. Deeg, K. Diercksen and C. Bräuchle, Ber. Bunsenges. Phys. Chem. 97, 1312 (1993).

    Google Scholar 

  137. P.G. de Gennes, in Liquid Crystals of One- and Two-Dimensionel Order, edited by W. Helfrich and G. Heppke, Springer, Berlin, p.231 (1980).

    Google Scholar 

  138. S. Götz, H. Scheuermann, W. Stille, and G. Strobl, Macromol. 26, 1529 (1993).

    Google Scholar 

  139. In order to use the equations of Ref. 36 (permanent network) for the polymer case (transient network), one has to add strain relaxation, i.e., (1/τ) ijkl Ψ kl to the strain quasi-current X ij in Eq. (4.23) of Ref. 36.

    Google Scholar 

  140. S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, Dover, New York (1981).

    Google Scholar 

  141. In the case of nematics this type of nonlinearity could also be classified as type ii, i.e., hidden in the director dependence of the material tensor λ ijk .

    Google Scholar 

  142. The use of ∇ i ρ, etc. besides ρ, etc. is necessary when describing inhomogeneous fluids, e.g., nematics or cholesterics with impurities, compare to Ref. 61.

    Google Scholar 

  143. Because of the higher order gradient terms, partial derivatives have to be replaced e.g., by ∂/(∂ρ) - ∇ i ,[∂/∂∇ i ρ)] +... or ∂/( i ,T) - ∇ j [∂/(∂∇ j i T)] +... when deriving thermodynamic conjugates or dissipative currents from the energy density and the entropy production, respectively.

    Google Scholar 

  144. H.R. Brand and H. Pleiner, Europhys. Lett. 26, 395 (1994).

    ADS  Google Scholar 

  145. L.E. Reichl, A Modern Course in Statistical Physics, Texas University, Austin (1980).

    Google Scholar 

  146. In a nonlinear description layer undulations also cause layer compression or dilation and k i i u A has to be replaced by k i i u A - (1/2)(δ ij - k i k j )(∇ i u A )(∇ j u A ), [compare to M. Kléman, Points, Lines and Walls, Wiley, New York (1983)].

    Google Scholar 

  147. F. Jähnig, J. Phys. (Paris) 36, 315 (1975).

    Google Scholar 

  148. R.E. Peierls, Ann. Inst. Henri Poincaré 5, 177 (1935) and

    MathSciNet  MATH  Google Scholar 

  149. L.D. Landau, Z. Sowjetunion II, 26 (1937).

    Google Scholar 

  150. G. Grinstein and R.A. Pelcovits, Phys. Rev. Lett. 47, 856 (1981).

    ADS  Google Scholar 

  151. G.F. Mazenko, S. Ramaswamy and T. Toner, Phys. Rev. Lett. 49, 51 (1982) and

    ADS  Google Scholar 

  152. S.T. Milner and P.C. Martin, Phys. Rev. Lett. 56, 77 (1986).

    ADS  Google Scholar 

  153. W. Helfrich, Phys. Rev. Lett. 23, 371 (1969).

    ADS  Google Scholar 

  154. P.G. de Gennes, Phys. Fluids 17, 1645 (1974).

    ADS  MATH  Google Scholar 

  155. W. Helfrich, Appl. Phys. Lett. 17, 531 (1970) and

    ADS  Google Scholar 

  156. J.P. Hurault, J. Chem. Phys. 59, 2068 (1973).

    ADS  Google Scholar 

  157. The sole difference being a reversible dynamic coupling between u and the vorticity parallel to the helix axis [7], which is absent in smectic systems, where the broken translational symmetry due to the layers is not related to any rotation about the layer axis.

    Google Scholar 

  158. ~ 3nm layer thickness in smectic systems versus ~ 0.5µ to ∞ pitch in cholesterics. This implies, for example, that the elastic constant for layer compression is much larger in smectic than in cholesteric liquid crystals [45].

    Google Scholar 

  159. In the Frank gradient energy this leads to terms linear in n curl n reflecting the lack of inversion symmetry of the helical structure.

    Google Scholar 

  160. O. Lehmann, Ann. Phys. (Leipzig) 2, 649 (1900).

    ADS  Google Scholar 

  161. H. Pleiner and H.R. Brand, J. Phys. Lett. (Paris) 41, L 383 (1980);

    MathSciNet  Google Scholar 

  162. H. Pleiner and H.R. Brand, Phys. Rev. A23, 944(1981).

    ADS  Google Scholar 

  163. H. Pleiner and H.R. Brand, Phys. Rev. A32, 3842 (1985).

    ADS  Google Scholar 

  164. Changes of the actual layer thickness (pitch length) due to a temperature gradient are of static (thermo-elastic) and dynamic (thermo-permeative) nature; compare to H. Pleiner and H.R. Brand, Mol. Cry st. Liq. Cryst. Lett. 2, 167 (1985).

    Google Scholar 

  165. A. Wulf, J. Chem. Phys. 59, 6596 (1973).

    ADS  Google Scholar 

  166. H. Pleiner and H.R. Brand, Mol. Cryst. Liq. Cryst. Lett. 7, 153 (1990).

    Google Scholar 

  167. H. Pleiner and H.R. Brand, J. Phys. II 3, 1397 (1993).

    Google Scholar 

  168. H.R. Brand, Makromol. Chem. Rap. Commun. 10, 147 (1989).

    ADS  Google Scholar 

  169. H.R. Brand and H. Pleiner, Makrokol. Chem. Rap. Commun. 11, 607 (1990).

    Google Scholar 

  170. W. Meier and H. Finkelmann, Makromol. Chem. Rap. Commun. 11, 599 (1990);

    Google Scholar 

  171. W. Meier and H. Finkelmann, Macromol. 26, 1811 (1993).

    ADS  Google Scholar 

  172. H. Pleiner and H.R. Brand, Mol. Cryst. Liq. Cryst. Lett. 5, 61 (1987).

    Google Scholar 

  173. H.R. Brand and H. Pleiner, Phys. Rev. A46, 3004 (1992).

    ADS  Google Scholar 

  174. Representing n in a spherical polar coordinate system with k the polar axis, the polar angle is the nonhydrodynamic and the azimuthal angle the hydrodynamic degree of freedom. In smectic A liquid crystals, where nk, both possible rotations of n are nonhydrodynamic.

    Google Scholar 

  175. For a linearized theory such a scenario seems to be reasonable, since layer compression or dilation costs much more energy than layer undulations. In the nonlinear domain, however, layer undulations are inevitably connected with changes of the layer thickness, (compare to Ref. 139).

    Google Scholar 

  176. F.M. Leslie, T. Carlsson, and N.A. Clark, Invited Talk at ILCC 15, Budapest (1994).

    Google Scholar 

  177. R.B. Meyer, L. Liébert, L. Strzelecki, and P. Keller, J. Phys. Lett. (Paris) 36, L 69 (1975).

    Google Scholar 

  178. H.R. Brand, P.E. Cladis, and P.L. Finn, Phys. Rev. A31, 361 (1985).

    ADS  Google Scholar 

  179. A. Schönfeld, F. Kremer, and R. Zentel, Liq. Cryst. 13, 403 (1993).

    Google Scholar 

  180. In-plane rotations together with c (or n) are the hydrodynamic (Goldstone) mode, describing rotations of the helix about its axis.

    Google Scholar 

  181. H. Pleiner and H.R. Brand, Ferroelectrics 148, 271 (1993).

    Google Scholar 

  182. H. Pleiner and H.R. Brand, Europhys. Lett. 28, 579 (1994).

    ADS  Google Scholar 

  183. H. Leube and H. Finkelmann, Makromol. Chem. 191, 2707 (1990);

    Google Scholar 

  184. H. Leube and H. Finkelmann, Makromol. Chem. 192, 1317 (1991).

    Google Scholar 

  185. H.R. Brand and H. Pleiner, Makromol Chem. Rap. Commun. 12, 539 (1991).

    Google Scholar 

  186. If these position vectors have equal lengths additionally, a perfect crystal is obtained.

    Google Scholar 

  187. D.R. Nelson and B.I Halperin, Phys. Rev. B21, 5312 (1980).

    ADS  Google Scholar 

  188. J.W. Goodby, Mol. Cryst. Liq. Cryst. 72, 95 (1980).

    Google Scholar 

  189. J.J. Benattar, F. Moussa, and M. Lambert, J. Chim. Phys. 80, 99 (1983).

    Google Scholar 

  190. In smectic F phases the tilt direction c coincides with one of the bond orientations in equilibrium (relative angle 0 mod π/3) and in the smectic I phase this relative angle is 30 degrees (mod π /3). For the hydrodynamic description this difference is irrelevant.

    Google Scholar 

  191. H.R. Brand and H. Pleiner, Phys. Rev. Lett. 59, 2822 (1987).

    ADS  Google Scholar 

  192. G.S. Smith, E.B. Sirota, C.R. Safinya and N.A. Clark, Phys. Rev. Lett. 60, 813 (1988).

    ADS  Google Scholar 

  193. J.V. Selinger and D.R. Nelson, Phys. Rev. Lett. 61, 416 (1988).

    MathSciNet  ADS  Google Scholar 

  194. D. Forster, T.C. Lubensky, P.C. Martin, J. Swift, and P.S. Pershan, Phys. Rev. Lett. 26, 1016 (1971).

    ADS  Google Scholar 

  195. In ref. [35] a slightly different representation of the viscosity tensor is given using transverse Kronecker tensors instead of the isotropic ones.

    Google Scholar 

  196. Arriving at Eq. (A. 16), as well as at Eqs. (2.59) and (A.2), we have made use of the electrostatic condition (2.50), i.e., ∇i E j = ∇i E i .

    Google Scholar 

  197. Due to our definition of the stress tensor in Eq. (A.2) σ ij in Eq. (A.7) has to be compared with — σ ij in the Leslie-Ericksen description.

    Google Scholar 

  198. D. Forster, Ann. Phys. (NY) 85, 505 (1974).

    ADS  Google Scholar 

  199. Instead of the formal relation λ = 0 one could also postulate div h = 0 to be a genuine part of “incompressibility” in nematics, but this also poses further restrictions on n i , and other quantities.

    Google Scholar 

  200. In simple liquids with \( {\sigma_{{ij}}} - (/2v)({\nabla_j}{v_i} + {\nabla_i}{v_j} - (2/3){\delta_{{ij}}} div\,v) - \zeta \,{\delta_{{ij}}}\,div\,v \) the formal relation 3ζ = v just eliminates Akk from σij and from ∇ j σ ij without changing any other aspect of the dynamics, since incompressibility (A kk = 0 and ρ = const.) is a special solution of the hydrodynamic equations (if thermal expansion is neglected or if T = const, is assumed additionally).

    Google Scholar 

  201. Classical elasticity theory [compare to L.D. Landau and E.M. Lifshitz, Theory of Elasticity, Pergamon, New York (1986) Section 22 and 23] shows that for isotropic systems longitudinal sound (connected to A kk ) is independent from transverse sound in the bulk (there is only a coupling via the surfaces for certain boundary conditions), and is therefore a true solution of the bulk linear elastodynamic equations, while for crystals with lower symmetry this is generally not true, thus rendering any incompressibility assumption unphysical.

    Google Scholar 

  202. If one considers only (∂/∂t) curl v three viscosities are sufficient (e.g., v1.2.3) and both Eqs. (A.23a) or (A.23b) would give the same result. However, the evaluation of the pressure via Eq. (A.26) requires one additional independent linear combination of viscosities (e.g., v 5 - v 4 + v 2). Thus four viscosities are needed to describe nematodynamics in the incompressibility approximation, while the Leslie-Ericksen approach contains only three, missing the δ ij n k n l A kl contribution to the pressure.

    Google Scholar 

  203. These susceptibilities are related to those introduced in Section 2.3 and used in Eqs. (A.27)-(A.29) by -α p = (K T /α s T)(Cv + σ α s T) ≈ (K T C V s T) and (1/K t ) = (l/K s ,) - (Cv/α 2s T) ≈ (C V /C p K s ) with C p = C V + (α 2p T/K T ).

    Google Scholar 

  204. S. Chandrasekhar, Liquid Crystals, Cambridge University, Cambridge (1977).

    Google Scholar 

  205. L.M. Blinov, Electro-optical and Magneto-optical Properties of Liquid Crystals, John Wiley, New York (1983).

    Google Scholar 

  206. L.M. Blinov and V.G. Chigrinov, Electrooptic Effects in Liquid Crystal Materials, Springer, New York (1994).

    Google Scholar 

  207. W. Urbach, Thèse d’Etat, Université Paris-Sud (1981) and

    Google Scholar 

  208. G. Ahlers, D.S. Cannell, L.I. Berge and S. Sakurai, Phys. Rev. E49, 545 (1994).

    Google Scholar 

  209. [1 D. Schmidt, M. Schadt and W. Helfrich, Z. Naturf. A27, 277 (1972).

    Google Scholar 

  210. J. Prost and P.S. Pershan, J. Appl. Phys. 47, 2298 (1976).

    ADS  Google Scholar 

  211. J.P. Marcerou and J. Prost, Ann. Phys. 3, 269 (1978).

    Google Scholar 

  212. M.I. Barnik, L.M. Blinov, A.N. Trufanov and B.A. Umanskii, Sov. Phys. JETP 46, 1016 (1977).

    ADS  Google Scholar 

  213. Z. Guozhen, J. Yuan, and S. Jianben, Phys. Lett. A109, 279 (1985).

    ADS  Google Scholar 

  214. L.M. Blinov, G. Durand, and S.V. Yablonsky, J. Phys. (Paris) II2, 1287 (1992).

    Google Scholar 

  215. B. Valenti, C. Bertoni, G. Barbero, P. Taverna-Valabrega, and R. Bartolino, Mol. Cry st. Liq. Cryst. 146, 307 (1987).

    Google Scholar 

  216. H. Pleiner, H.R. Brand and W. Zimmermann, to be published.

    Google Scholar 

  217. J.D. Jackson, Classical Electrodynamics, 2nd ed., Wiley, New York (1975).

    MATH  Google Scholar 

Download references

Authors

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 1996 Springer-Verlag New York, Inc.

About this chapter

Cite this chapter

Pleiner, H., Brand, H.R. (1996). Hydrodynamics and Electrohydrodynamics of Liquid Crystals. In: Buka, A., Kramer, L. (eds) Pattern Formation in Liquid Crystals. Partially Ordered Systems. Springer, New York, NY. https://doi.org/10.1007/978-1-4612-3994-9_2

Download citation

  • DOI: https://doi.org/10.1007/978-1-4612-3994-9_2

  • Publisher Name: Springer, New York, NY

  • Print ISBN: 978-1-4612-8464-2

  • Online ISBN: 978-1-4612-3994-9

  • eBook Packages: Springer Book Archive

Publish with us

Policies and ethics