Skip to main content
Log in

On the Landau–de Gennes Elastic Energy of a Q-Tensor Model for Soft Biaxial Nematics

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

In the Landau–de Gennes theory of liquid crystals, the propensities for alignments of molecules are represented at each point of the fluid by an element \(\mathbf{Q}\) of the vector space \({\mathcal {S}}_0\) of \(3\times 3\) real symmetric traceless matrices, or \(\mathbf{Q}\)-tensors. According to Longa and Trebin (1989), a biaxial nematic system is called soft biaxial if the tensor order parameter \(\mathbf{Q}\) satisfies the constraint \(\mathrm{tr}(\mathbf{Q}^2) = \text {const}\). After the introduction of a \(\mathbf{Q}\)-tensor model for soft biaxial nematic systems and the description of its geometric structure, we address the question of coercivity for the most common four-elastic-constant form of the Landau–de Gennes elastic free-energy (Iyer et al. 2015) in this model. For a soft biaxial nematic system, the tensor field \(\mathbf{Q}\) takes values in a four-dimensional sphere \({{\mathbb {S}}}^4_\rho \) of radius \(\rho \le \sqrt{2/3}\) in the five-dimensional space \({\mathcal {S}}_0\) with inner product \(\langle \mathbf{Q}, {\mathbf {P}} \rangle = \mathrm{tr}(\mathbf{Q}{\mathbf {P}})\). The rotation group \(\textit{SO}(3)\) acts orthogonally on \({\mathcal {S}}_0\) by conjugation and hence induces an action on \({{\mathbb {S}}}^4_\rho \subset {\mathcal {S}}_0\). This action has generic orbits of codimension one that are diffeomorphic to an eightfold quotient \({{\mathbb {S}}}^3/{\mathcal {H}}\) of the unit three-sphere \({{\mathbb {S}}}^3\), where \({{\mathcal {H}}}=\{\pm 1, \pm \mathsf{i}, \pm \mathsf{j}, \pm \mathsf{k}\}\) is the quaternion group, and has two degenerate orbits of codimension two that are diffeomorphic to the projective plane \({\mathbb {R}}P^2\). Each generic orbit can be interpreted as the order parameter space of a constrained biaxial nematic system and each singular orbit as the order parameter space of a constrained uniaxial nematic system. It turns out that \({{\mathbb {S}}}^4_\rho \) is a cohomogeneity one manifold, i.e., a manifold with a group action whose orbit space is one-dimensional. Another important geometric feature of the model is that the set \(\Sigma _\rho \) of diagonal \(\mathbf{Q}\)-tensors of fixed norm \(\rho \) is a (geodesic) great circle in \({{\mathbb {S}}}^4_\rho \) which meets every orbit of \({{\mathbb {S}}}^4_\rho \) orthogonally and is then a section for \({{\mathbb {S}}}^4_\rho \) in the sense of the general theory of canonical forms. We compute necessary and sufficient coercivity conditions for the elastic energy by exploiting the \(\textit{SO}(3)\)-invariance of the elastic energy (frame-indifference), the existence of the section \(\Sigma _\rho \) for \({{\mathbb {S}}}^4_\rho \), and the geometry of the model, which allow us to reduce to a suitable invariant problem on (an arc of) \(\Sigma _\rho \). Our approach can ultimately be seen as an application of the general method of reduction of variables, or cohomogeneity method.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. Observe that this conditions holds on an open and dense subset.

  2. Note that \({\varvec{\Lambda }}(-t)\) is obtained from \({\varvec{\Lambda }}(t)\) by interchanging the first two eigenvalues.

References

  • Alekseevsky, A.V., Alekseevsky, D.V.: Riemannian \(G\)-manifold with one-dimensional orbit space. Ann. Glob. Anal. Geom. 11(3), 197–211 (1993)

    MathSciNet  MATH  Google Scholar 

  • Allender, D., Longa, L.: Landau-de Gennes theory of biaxial nematics reexamined. Phys. Rev. E 78, 011704–011714 (2008)

    Article  Google Scholar 

  • Ball, J.M.: The Mathematics of Liquid Crystals. Cambridge Centre for Analysis Short Course. https://people.maths.ox.ac.uk/ball/teaching.shtml. 13–17 Feb 2012

  • Ball, J.M., Majumdar, A.: Nematic liquid crystals: from Maier–Saupe to a continuum theory. Mol. Cryst. Liq. Cryst. 525, 1–11 (2010)

    Article  Google Scholar 

  • Ball, J.M., Zarnescu, A.: Orientability and energy minimization in liquid cristal models. Arch. Ration. Mech. Anal. 202, 493–535 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Berreman, D.W., Meiboom, S.: Tensor representation of Oseen–Frank strain energy in uniaxial cholesterics. Phys. Rev. A 30, 1955–1959 (1984)

    Article  Google Scholar 

  • Bredon, G.E.: Introduction to Compact Transformation Groups. Pure and Applied Mathematics, vol. 46. Academic Press, New York (1972)

    MATH  Google Scholar 

  • Cartan, E.: Sur des families remarquables d’hypersurfaces isoparamétriques dans les espaces sphériques. Math. Z. 45, 335–367 (1939)

    Article  MathSciNet  MATH  Google Scholar 

  • Chillingworth, D.: Perturbed hedgehogs: continuous deformation of point defects in biaxial nematic liquid crystals. IMA J. Appl. Math. 81(4), 647–661 (2016)

    Article  MathSciNet  Google Scholar 

  • Davis, T.A., Gartland Jr., E.C.: Finite element analysis of the Landau–de Gennes minimization problem for liquid crystals. SIAM J. Numer. Anal. 35, 336–362 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  • de Gennes, P.G., Prost, J.: The Physics of Liquid Crystals. International Series of Monographs on Physics, vol. 83, 2nd edn. Oxford University Press, Oxford (1993)

    Google Scholar 

  • Ericksen, J.L.: Inequalities in liquid crystal theory. Phys. Fluids 9, 1205–1207 (1966)

    Article  Google Scholar 

  • Ericksen, J.L., Kinderlehrer, D. (eds.): Theory and Applications of Liquid Crystals. IMA Volumes in Mathematics and Its Applications, vol. 5. Springer, New York (1987)

    Google Scholar 

  • Frank, F.C.: I. Liquid crystals. On the theory of liquid crystals. Discuss. Faraday Soc. 25, 19–28 (1958)

    Article  Google Scholar 

  • Govers, E., Vertogen, G.: Elastic continuum theory of biaxial nematics. Phys. Rev. A 30, 1998–2000 (1984)

    Article  Google Scholar 

  • Govers, E., Vertogen, G.: Elastic continuum theory of biaxial nematics. Phys. Rev. A 31, 1957 (1985)

    Article  Google Scholar 

  • Gramsbergen, E.F., Longa, L., de Jeu, W.H.: Landau theory of the nematic–isotropic phase transition. Phys. Rep. 135, 195–257 (1986)

    Article  Google Scholar 

  • Hardt, R., Kinderlehrer, D.: Mathematical questions of liquid crystal theory. In: Ericksen, J.L., Kinderlehrer, D. (eds.) Theory and Applications of Liquid Crystals. IMA Volumes in Mathematics and Its Applications, vol. 5, pp. 151–184. Springer, New York (1987)

    Chapter  Google Scholar 

  • Hsiang, W.-Y.: On the compact homogeneous minimal submanifolds. Proc. Natl. Acad. Sci. USA 56, 5–6 (1966)

    Article  MathSciNet  MATH  Google Scholar 

  • Hsiang, W.-Y., Lawson Jr., H.B.: Minimal submanifolds of low cohomogeneity. J. Differ. Geom. 5, 1–38 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  • Iyer, G., Xu, X., Zarnescu, A.: Dynamic cubic instability in a 2D Q-tensor model for liquid crystals. Math. Models Methods Appl. Sci. 25, 1477–1517 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Landau, L.D., Lifshitz, E.M.: Statistical Physics, 2nd edn. Pergamon, Oxford (1969)

    MATH  Google Scholar 

  • Lawson Jr., H.B.: Lectures on Minimal Submanifolds. Mathematics Lecture Series, 9, vol. I, 2nd edn. Publish or Perish, Inc., Wilmington (1980)

    Google Scholar 

  • Lin, F.H., Liu, C.: Static and dynamic theories of liquid crystals. J. Partial Differ. Equ. 14, 289–330 (2001)

    MathSciNet  MATH  Google Scholar 

  • Longa, L.: Private communication, March, 2016

  • Longa, L., Trebin, H.-R.: Structure of the elastic free energy for chiral nematic liquid crystals. Phys. Rev. A 39, 2160–2168 (1989)

    Article  MathSciNet  Google Scholar 

  • Longa, L., Monselesan, D., Trebin, H.-R.: An extension of the Landau–Ginzburg–de Gennes theory for liquid crystals. Liq. Cryst. 2, 769–796 (1987)

    Article  Google Scholar 

  • Longa, L., Pajak, G., Wydro, T.: Stability of biaxial nematic phase for systems with variable molecular shape anisotropy. Phys. Rev. E 76, 011703–011709 (2007)

    Article  Google Scholar 

  • Majumdar, A.: Equilibrium order parameters of nematic liquid crystals in the Landau–de Gennes theory. Eur. J. Appl. Math. 21, 181–203 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Mermin, N.D.: The topological theory of defects in ordered media. Rev. Mod. Phys. 51, 591–648 (1979)

    Article  MathSciNet  Google Scholar 

  • Messiah, A.: Quantum Mechanics. Dover Publications Inc., Mineola (2003)

    MATH  Google Scholar 

  • Michel, L.: Points critiques des fonctions invariantes sur une \(G\)-variété, C. R. Acad. Sci. Paris Sér. A-B 272, A433–A436 (1971)

    Google Scholar 

  • Michel, L.: Symmetry defects and broken symmetry. Configurations hidden symmetry. Rev. Mod. Phys. 52(3), 617–651 (1980)

    Article  MathSciNet  Google Scholar 

  • Mkaddem, S., Gartland Jr., E.C.: Fine structure of defects in radial nematic droplets. Phys. Rev. E 62, 6694–6705 (2000)

    Article  Google Scholar 

  • Mostert, P.: On a compact Lie group acting on a manifold. Ann. Math., 65, pp. 447–455, (1957); Errata, Ann. of Math., 66 (1957), 589

  • Mottram, N.J., Newton, C.: Introduction to \(Q\)-tensor theory, Technical report, University of Strathclyde, Department of Mathematics (2004). arXiv:1409.3542 [cond-mat.soft], 2014

  • Mucci, D.: Maps into projective spaces: liquid crystals and conformal energies. Discrete Contin. Dyn. Syst. Ser. B 17, 597–635 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Mucci, D., Nicolodi, L.: On the elastic energy density of constrained Q-tensor models for biaxial nematics. Arch. Ration. Mech. Anal. 206, 853–884 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Mucci, D., Nicolodi, L.: On the Landau–de Gennes elastic energy of constrained biaxial nematics. SIAM J. Math. Anal. 48(3), 1954–1987 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Nicolodi, L.: Models for biaxial nematic liquid crystals. J. Phys. Conf. Ser. 410, 012043–012046 (2013)

    Article  Google Scholar 

  • O’Neill, B.: Semi-Riemannian Geometry. With Application to Relativity, Pure and Applied Mathematics, vol. 103. Academic Press, Inc., New York (1983)

    Google Scholar 

  • Oseen, C.W.: The theory of liquid crystals. Trans. Faraday Soc. 29, 883–899 (1933)

    Article  MATH  Google Scholar 

  • Palais, R.S., Terng, C.-L.: A general theory of canonical forms. Trans. Am. Math. Soc. 300, 771–789 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  • Palais, R.S., Terng, C.-L.: Critical Point Theory and Submanifold Geometry, Lecture Notes in Mathematics, vol. 1353. Springer, Berlin (1988)

  • Penzenstadler, E., Trebin, H.-R.: Fine structure of point defects and soliton decay in nematic liquid crystals. J. Phys. Fr. 50, 1027–1040 (1989)

    Article  Google Scholar 

  • Rosso, R., Virga, E.G.: Metastable nematic hedgehogs. J. Phys. A 29, 4247–4264 (1996)

    Article  MATH  Google Scholar 

  • Schopohl, N., Sluckin, T.J.: Defect core structure in nematic liquid crystals. Phys. Rev. Lett. 59, 2582–2584 (1987)

    Article  Google Scholar 

  • Sonnet, A., Kilian, A., Hess, S.: Alignment tensor versus director: description of defects in nematic liquid crystals. Phys. Rev. E 52, 718–722 (1995)

    Article  Google Scholar 

  • Takagi, R., Takahashi, T.: On the principal curvatures of homogeneous hypersurfaces in a sphere. In: Differential Geometry in Honor of Kentaro Yano, pp. 469–481. Kinokuniya, Tokyo (1972)

  • Virga, E.: Variational Theories for Liquid Crystals, Applied Mathematics and Mathematical Computation, vol. 8. Chapman & Hall, London (1994)

    Book  MATH  Google Scholar 

  • Willmore, T.J.: Total Curvature in Riemannian Geometry. Hellis Horwood Series in Mathematics and its Applications. Hellis Horwood, Chichester (1982)

    Google Scholar 

  • Zocher, H.: The effect of a magnetic field on the nematic state. Trans. Faraday Soc. 29, 945–957 (1933)

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank Professor Lech Longa for his kind interest. The authors would also like to thank the anonymous referees for their useful comments and suggestions. This research was partially supported by PRIN 2010–2011 “Calcolo delle Variazioni” (D.M.) and by PRIN 2010-2011 “Varietà reali e complesse: geometria, topologia e analisi armonica” (L.N.). The authors are members of the “Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni” (GNAMPA) (D.M.) and of the “Gruppo Nazionale per le Strutture Algebriche, Geometriche e le loro Applicazioni” (GNSAGA) (L.N.) of the Istituto Nazionale di Alta Matematica “F. Severi” (INdAM).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lorenzo Nicolodi.

Additional information

Communicated by Irene Fonseca.

Appendix: The Longa–Monselesan–Trebin Conditions

Appendix: The Longa–Monselesan–Trebin Conditions

In this appendix, using the scalar coordinates corresponding to the representation of \(\mathbf{Q}\)-tensors described in Sect. 2.1, we compute the Longa–Monselesan–Trebin positivity conditions

$$\begin{aligned} 2L_3>L_2,\quad L_3+L_2>0,\quad 10L_1+L_2+6L_3>0 \end{aligned}$$
(5.1)

for the three-elastic-constant form \(L_1I_1+L_2I_2+L_3I_3\) of the elastic energy density \(\psi _E\) in the general biaxial case (cf. (1.8)).

These conditions were originally obtained by Longa–Monselesan–Trebin (1987) (cf. also Davis and Gartland 1998) by writing the elastic energy density \(L_1I_1+L_2I_2+L_3I_3\) as a linear combination of irreducible \(\textit{SO}(3)\)-invariants, computed using the representation theory of \(\textit{SO}(3)\) on spherical tensors and the Clebsch–Gordan coefficients from the angular momentum theory of quantum mechanics (Messiah 2003).

In general, by (2.7) we have \(I_3=|\nabla \mathbf{u}|^2\). Also, by (3.7), in terms of \(\mathbf{u}\) we have

$$\begin{aligned} 2I_1= & {} \left( {\frac{1}{\sqrt{3}}}\partial _1u^1+\partial _1u^2+\partial _2u^3+\partial _3u_4\right) ^2+ \left( {\frac{1}{\sqrt{3}}}\partial _2u^1-\partial _2u^2+ \partial _3u^5+\partial _1u^3\right) ^2\\&+\left( -{\frac{2}{\sqrt{3}}}\partial _3u^1 +\partial _1u^4+\partial _2u^5\right) ^2. \end{aligned}$$

Similarly, by (3.8) we find the formula

$$\begin{aligned} 2I_2= & {} \displaystyle \left( {\frac{1}{\sqrt{3}}}\partial _1u^1+\partial _1u^2\right) ^2+(\partial _2u^3)^2+(\partial _3u^4)^2\\&+\,2\Bigl (-{\frac{2}{\sqrt{3}}}\partial _1u^1 \partial _3u^4+{\frac{1}{\sqrt{3}}}\partial _1u^1\partial _2u^3 - \partial _1u^2\partial _2u^3\Bigr )\\&+\,\displaystyle \left( {\frac{1}{\sqrt{3}}}\partial _2u^1-\partial _2u^2\right) ^2+(\partial _3u^5)^2+(\partial _1u^3)^2\\&+\,2\left( -{\frac{2}{\sqrt{3}}}\partial _2u^1\partial _3u^5+{\frac{1}{\sqrt{3}}}\partial _2u^1\partial _1u^3 - \partial _2u^2\partial _1u^3\right) \\&+\,\displaystyle \Bigl (-{\frac{2}{\sqrt{3}}}\partial _3u^1\Bigr )^2+(\partial _1u^4)^2+(\partial _2u^5)^2\\&+\,2\left( {\frac{1}{\sqrt{3}}}\partial _3u^1\partial _1u^4 + \partial _3u^2\partial _1u^4+ {\frac{1}{\sqrt{3}}}\partial _3u^1\partial _2u^5 - \partial _3u^2\partial _2u^5\right) \\&+\,2\,\bigl (\partial _3u^3\partial _2u^4+\partial _3u^3\partial _1u^5+\partial _2u^4\partial _1u^5\bigr ) . \end{aligned}$$

For any choice of the real coefficients \(L_i\) we thus may decompose into four quadratic forms:

$$\begin{aligned} L_1I_1+L_2I_2+L_3I_3={{\mathcal {F}}}_1+{{\mathcal {F}}}_2+{{\mathcal {F}}}_3+{{\mathcal {F}}}_4. \end{aligned}$$

The quadratic form \({{\mathcal {F}}}_1\) depends on the four variables \(\partial _1u^1,\,\partial _1u^2,\,\partial _2u^3,\,\partial _3u_4\), and its related matrix is:

$$\begin{aligned} M_1:=\begin{pmatrix} \displaystyle L_3+{\frac{L_1+L_2}{6}} &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(L_1-2L_2) \\ \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{K}{2}} &{}\quad \displaystyle {\frac{1}{2}}(L_1-L_2) &{}\quad \displaystyle {\frac{L_1}{2}} \\ \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{1}{2}}(L_1-L_2) &{}\quad \displaystyle {\frac{K}{2}} &{}\quad \displaystyle {\frac{L_1}{2}} \\ \displaystyle {\frac{1}{2\sqrt{3}}}(L_1-2L_2) &{}\quad \displaystyle {\frac{L_1}{2}} &{}\quad \displaystyle {\frac{L_1}{2}} &{}\quad \displaystyle {\frac{K}{2}} \end{pmatrix} \end{aligned}$$

where we have denoted \(K:=L_1+L_2+2L_3\). The second one, \({{\mathcal {F}}}_2\), depends on the four variables \(\partial _2u^1,\) \(\partial _2u^2,\) \(\partial _3u^5,\) \(\partial _1u^3\), and its corresponding matrix is:

$$\begin{aligned} M_2:=\begin{pmatrix} \displaystyle L_3+{\frac{L_1+L_2}{6}} &{}\quad \displaystyle -{\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(L_1-2L_2) &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) \\ \displaystyle -{\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle {\frac{K}{2}} &{}\quad \displaystyle -{\frac{L_1}{2}} &{}\quad -\displaystyle {\frac{1}{2}}(L_1-L_2) \\ \displaystyle {\frac{1}{2\sqrt{3}}}(L_1-2L_2) &{}\quad -\displaystyle {\frac{L_1}{2}} &{}\quad \displaystyle {\frac{K}{2}} &{}\quad \displaystyle {\frac{L_1}{2}} \\ \displaystyle {\frac{1}{2\sqrt{3}}}(L_1+L_2) &{}\quad \displaystyle -{\frac{1}{2}}(L_1-L_2) &{}\quad \displaystyle {\frac{L_1}{2}} &{}\quad \displaystyle {\frac{K}{2}} \end{pmatrix}. \end{aligned}$$

The third one, \({{\mathcal {F}}}_3\), depends on the four variables \(\partial _3u^1,\,\partial _3u^2,\,\partial _1u^4,\,\partial _2u^5\), and its corresponding matrix is:

$$\begin{aligned} M_3:=\begin{pmatrix} \displaystyle L_3+{\frac{2}{3}}(L_1+L_2) &{}\quad \displaystyle 0 &{}\quad \displaystyle {\frac{1}{2\sqrt{3}}}(-2L_1+L_2) &{} \displaystyle {\frac{1}{2\sqrt{3}}}(-2L_1+L_2) \\ \displaystyle 0 &{}\quad \displaystyle L_3 &{}\quad \displaystyle {\frac{L_2}{2}} &{}\quad -\displaystyle {\frac{L_2}{2}} \\ \displaystyle {\frac{1}{2\sqrt{3}}}(-2L_1+L_2) &{}\quad \displaystyle {\frac{L_2}{2}}&{}\quad \displaystyle {\frac{K}{2}} &{} \quad \displaystyle {\frac{L_1}{2}} \\ \displaystyle {\frac{1}{2\sqrt{3}}}(-2L_1+L_2) &{}\quad \displaystyle -{\frac{L_2}{2}} &{}\quad \displaystyle {\frac{L_1}{2}} &{}\quad \displaystyle {\frac{K}{2}} \end{pmatrix}. \end{aligned}$$

Finally, the fourth one, \({{\mathcal {F}}}_4\), depends on the three remaining variables \(\partial _3u^3,\,\partial _2u^4,\,\partial _1u^5\), and its corresponding matrix is:

$$\begin{aligned} M_4:=\begin{pmatrix} \displaystyle L_3 &{} \displaystyle {\frac{L_2}{2}} &{}\quad \displaystyle {\frac{L_2}{2}} \\ \displaystyle \displaystyle {\frac{L_2}{2}} &{}\quad \displaystyle L_3 &{}\quad \displaystyle {\frac{L_2}{2}} \\ \displaystyle {\frac{L_2}{2}}&{}\quad \displaystyle {\frac{L_2}{2}} &{}\quad L_3 \end{pmatrix}. \end{aligned}$$

The matrix \(M_4\) has determinant \((L_3-L_2/2)^2(L_3+L_2)\), whence \({{\mathcal {F}}}_4\) is positive definite if and only if the conditions \(L_3>0\), \(2L_3>|L_2|\), \(L_3+L_2>0\) hold, which clearly reduce to the system

$$\begin{aligned} L_3>0,\quad 2L_3>L_2,\quad L_3+L_2>0. \end{aligned}$$
(5.2)

Dealing with \(M_3\), and starting from the right-bottom corner, we obtain the first two conditions \(K>0\) and \(K>|L_1|\), which reduce to

$$\begin{aligned} L_2+2L_3>0,\quad 2L_1+L_2+2L_3>0, \end{aligned}$$
(5.3)

where the first inequality follows from (5.2). The determinant of the \(3\times 3\) right-bottom minor is

$$\begin{aligned}&{\frac{1}{4}}\,(2L_1+L_2+2L_3)(2L_3^2+L_2L_3-L_2^2)\\&\quad ={\frac{1}{4}}\,(2L_1+L_2+2L_3)(2L_3-L_2)(L_3+L_2) \end{aligned}$$

and hence it is positive under the assumptions (5.2) and (5.3). Finally, we compute \(\mathrm{det}\,M_3\) by applying Laplace’s formula w.r.t. the first column, and we write

$$\begin{aligned} \mathrm{det}\,M_3=A_3^1-A_3^2+A_3^3-A_3^4 \end{aligned}$$

where \(A_3^2=0\) and we, respectively, compute:

$$\begin{aligned} A_3^1= & {} \left( L_3+{\frac{2}{3}}\,(L_1+L_2) \right) \cdot {\frac{1}{4}}\,(2L_1+L_2+2L_3)(2L_3-L_2)(L_3+L_2),\\ A_3^3-A_3^4= & {} -{\frac{1}{12}}\,(-2L_1+L_2)^2\,(2L_3-L_2)(L_3+L_2) \end{aligned}$$

and hence

$$\begin{aligned} \begin{array}{rl} 12\mathrm{det}\,M_3= &{} (2L_3-L_2)(L_3+L_2)\,[(3L_3+2L_1+2L_2)(2L_1+L_2+2L_3)\\ &{}\quad -(-2L_1+L_2)^2] \\ =&{} \displaystyle (2L_3-L_2)(L_3+L_2)^2(6L_3+L_1+10L_2). \end{array} \end{aligned}$$

Therefore, under the conditions (5.2) and (5.3), the determinant of \(M_3\) is positive if and only if we also have

$$\begin{aligned} 6L_3+L_1+10L_2>0 . \end{aligned}$$
(5.4)

We now consider the matrix \(M_2\). Starting from the right-bottom corner, we again obtain the first two conditions \(K>0\) and \(K>|L_1|\). This time, the determinant D of the \(3\times 3\) right-bottom minor is

$$\begin{aligned} D= & {} {\frac{1}{8}}\,\bigl [K(K^2-2L_1^2-(L_2-L_1)^2)-2L_1^2(L_2-L_1) \bigr ]\nonumber \\= & {} {\frac{1}{2}}\,(L_3+L_2)(2L_3^2+(3L_1+L_2)L_3+L_1L_2) \end{aligned}$$
(5.5)

and hence it is positive under the additional assumption

$$\begin{aligned} 2L_3^2+(3L_1+L_2)L_3+L_1L_2>0. \end{aligned}$$
(5.6)

As before, we now compute \(\mathrm{det}\,M_2\) by applying Laplace’s formula w.r.t. the first column, and we write

$$\begin{aligned} \mathrm{det}\,M_2=A_2^1-A_2^2+A_2^3-A_2^4. \end{aligned}$$

We have:

$$\begin{aligned} A_2^1= & {} {\frac{6L_3+L_1+L_2}{6}}\cdot {\frac{1}{2}}\,(L_3+L_2)(2L_3^2+(3L_1+L_2)L_3+L_1L_2),\\ -A_2^2= & {} -{\frac{1}{24}}\,(L_3+L_2)(L_1+L_2)(4L_1L_2+L_2^2+2L_1L_3+2L_2L_3),\\ A_2^3= & {} {\frac{1}{24}}\,(L_3+L_2)(L_1-2L_2)(3L_1L_2-L_1L_3+2L_2L_3),\\ -A_2^4= & {} -{\frac{1}{24}}\,(L_3+L_2)(L_1+L_2)(4L_1L_2+L_2^2+2L_1L_3+2L_2L_3)=-A_2^2, \end{aligned}$$

and hence we obtain:

$$\begin{aligned} \begin{array}{rl} 12\,\mathrm{det}\,M_2= (L_2+L_3)\,\bigl [ &{} (6L_3+L_1+L_2)(2L_3^2+(3L_1+L_2)L_3+L_1L_2) \\ &{}-(L_1+L_2)(4L_1L_2+L_2^2+2L_1L_3+2L_2L_3) \\ &{} + (L_1-2L_2)(3L_1L_2-L_1L_3+2L_2L_3) \,\,\bigr ]\\ = \,\, (L_2+L_3)\,\bigl [ &{} 12L_3^3+4(5L_1+2L_2)L_3^2+(10L_1L_2-5L_2^2)L_3\\ &{}-(10L_1L_2^2+L_2^3) \,\,\bigr ] \\ = \,\, (L_2+L_3)\,\bigl [ &{} (L_2+L_2)(2L_3-L_2)(6L_3+10L_1+L_2)\,\,\bigr ]. \end{array} \end{aligned}$$

Therefore, condition \(\mathrm{det}\,M_3>0\) follows from (5.2), (5.3), and (5.4).

We finally deal with the matrix \(M_1\), obtaining exactly the same positivity conditions as for \(M_2\). In fact, starting from the right-bottom corner, the first two conditions are again \(K>0\) and \(K>|L_1|\), whereas the determinant D of the \(3\times 3\) right-bottom minor is equal to the expression from (5.5). Computing as before the determinant of \(M_1\) by means of Laplace’s formula w.r.t. the first column, and writing

$$\begin{aligned} \mathrm{det}\,M_1=A_1^1-A_1^2+A_1^3-A_1^4, \end{aligned}$$

it is not difficult to check that we have:

$$\begin{aligned} A_1^1=A_1^2,\quad -A_1^2=-A_2^2,\quad A_1^3=-A_2^4,\quad -A_1^4=A_2^3 \end{aligned}$$

and hence we get \(\mathrm{det}\,M_1=\mathrm{det}\,M_2\).

In conclusion, arguing as in Remark 3.4 we deduce that the system (5.2), (5.3), (5.4), and (5.6) is equivalent to the one in (5.1).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mucci, D., Nicolodi, L. On the Landau–de Gennes Elastic Energy of a Q-Tensor Model for Soft Biaxial Nematics. J Nonlinear Sci 27, 1687–1724 (2017). https://doi.org/10.1007/s00332-017-9383-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-017-9383-4

Keywords

Mathematics Subject Classification

Navigation