Skip to main content
Log in

Refined asymptotics for Landau-de Gennes minimizers on planar domains

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

In our previous work Golovaty and Montero (Arch Ration Mech Anal 213(2):447–490, 2014), we studied asymptotic behavior of minimizers of the Landau-de Gennes energy functional on planar domains as the nematic correlation length converges to zero. Here we improve upon those results, in particular by sharpening the description of the limiting map of the minimizers. We also provide an expression for the energy valid for a small, but fixed value of the nematic correlation length.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

References

  1. Baldo, S., Jerrard, R.L., Orlandi, G., Soner, H.M.: Convergence of Ginzburg-Landau functionals in 3-d superconductivity. Arch. Rat. Mech. Anal. 25, 699–752 (2012)

    Article  MathSciNet  Google Scholar 

  2. Bethuel, F., Brezis, H., Hélein, F.: Asymptotics for the minimization of a Ginzburg-Landau functional. Calculus of Variations and Partial Differential Equations 1(2), 123–148 (1993)

    Article  MathSciNet  Google Scholar 

  3. Bethuel, F., Brezis, H., Hélein, F.: Ginzburg-Landau vortices. Modern Birkhäuser Classics. Birkhäuser/Springer, Cham, (2017). Reprint of the 1994 edition [MR1269538]

  4. Canevari, G.: Biaxiality in the asymptotic analysis of a \(2\)-d Landau-de Gennes model for liquid crystals. ESIAM-COCV 21, 101–137 (2015)

    Article  MathSciNet  Google Scholar 

  5. Canevari, G.: Line defects in the small elastic constant limit of a three-dimensional Landau-de Gennes model. Arch. Ration. Mech. Anal. 223(2), 591–676 (2017)

    Article  MathSciNet  Google Scholar 

  6. Canevari, G., Zarnescu, A.: Design of effective bulk potentials for nematic liquid crystals via colloidal homogenisation. Math. Models Methods Appl. Sci. 30(2), 309–342 (2020)

    Article  MathSciNet  Google Scholar 

  7. Canevari, G., Zarnescu, A.: Polydispersity and surface energy strength in nematic colloids. Math. Eng. 2(2), 290–312 (2020)

    Article  MathSciNet  Google Scholar 

  8. COMSOL Multiphysics® v. 5.3. http://www.comsol.com/. COMSOL AB, Stockholm, Sweden

  9. Di Fratta, G., Robbins, J.M., Slastikov, V., Zarnescu, A.: Half-integer point defects in the \(Q\)-tensor theory of nematic liquid crystals. J. Nonlinear Sci. 26(1), 121–140 (2016)

    Article  MathSciNet  Google Scholar 

  10. Di Fratta, Giovanni, Robbins, Jonathan M., Slastikov, Valeriy, Zarnescu, Arghir: Landau-de Gennes corrections to the Oseen-Frank theory of nematic liquid crystals. Arch. Ration. Mech. Anal. 236(2), 1089–1125 (2020)

    Article  MathSciNet  Google Scholar 

  11. Dorfmeister, Josef, McIntosh, Ian, Pedit, Franz, Hongyou, Wu.: On the meromorphic potential for a harmonic surface in a \(k\)-symmetric space. Manuscripta Math. 92(2), 143–152 (1997)

    Article  MathSciNet  Google Scholar 

  12. Gilbarg, D., Trudinger, N.S.: Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, (2001). Reprint of the 1998 edition

  13. Golovaty, D., Montero, J.A.: On minimizers of a Landau-de Gennes energy functional on planar domains. Arch. Ration. Mech. Anal. 213(2), 447–490 (2014)

    Article  MathSciNet  Google Scholar 

  14. Hélein, F.: Harmonic Maps, Conservation Laws and Moving Frames. Cambridge Tracts in Mathematics. Cambridge University Press, (2004)

  15. Henao, D., Majumdar, A.: Symmetry of uniaxial global Landau-de Gennes minimizers in the theory of nematic liquid crystals. SIAM J. Math. Anal. 44(5), 3217–3241 (2012)

    Article  MathSciNet  Google Scholar 

  16. Henao, D., Majumdar, A., Pisante, A.: Uniaxial versus biaxial character of nematic equilibria in three dimensions. Calc. Var. Partial Differential Equations, 56(2):Paper No. 55, 22, (2017)

  17. Hitchin, N., Segal, D., Ward, R.: Integrable systems: Twistors. Loop Groups and Riemann Surfaces. Clarendon Press, Oxford (1999)

  18. Hörmander, L.: Linear partial differential operators. A series of comprehensive studies in mathematics. Springer-Verlag, (1977). Fourth printing of the 1963 edition

  19. Ignat, R., Nguyen, L., Slastikov, V., Zarnescu, A.: Stability of the melting hedgehog in the Landau-de Gennes theory of nematic liquid crystals. Arch. Ration. Mech. Anal. 215(2), 633–673 (2015)

    Article  MathSciNet  Google Scholar 

  20. Ignat, R., Nguyen, L., Slastikov, V., Zarnescu, A.: Instability of point defects in a two-dimensional nematic liquid crystal model. Ann. Inst. H. Poincaré Anal. Non Linéaire 33(4), 1131–1152 (2016)

    Article  MathSciNet  Google Scholar 

  21. Ignat, R., Nguyen, L., Slastikov, V., Zarnescu, A.: Stability of point defects of degree \(\pm \frac{1}{2}\) in a two-dimensional nematic liquid crystal model. Calc. Var. Partial Differential Equations, 55(5):Art. 119, 33, (2016)

  22. Ignat, R., Nguyen, L., Slastikov, V., Zarnescu, A.: Symmetry and multiplicity of solutions in a two-dimensional Landau-de Gennes model for liquid crystals. Arch. Ration. Mech. Anal. 237(3), 1421–1473 (2020)

    Article  MathSciNet  Google Scholar 

  23. Jerrard, R.L., Soner, H.M.: The Jacobian and the Ginzburg-Landau energy. Calc. Var and PDE 145, 151–191 (2002)

    Article  MathSciNet  Google Scholar 

  24. Kitavtsev, G., Robbins, J.M., Slastikov, V., Zarnescu, A.: Liquid crystal defects in the Landau–de Gennes theory in two dimensions—beyond the one-constant approximation. Math. Models Methods Appl. Sci. 26(14), 2769–2808 (2016)

    Article  MathSciNet  Google Scholar 

  25. Majumdar, A., Zarnescu, A.: Landau-de Gennes theory of nematic liquid crystals: the Oseen-Frank limit and beyond. Arch. Ration. Mech. Anal. 196(1), 227–280 (2010)

    Article  MathSciNet  Google Scholar 

  26. Monteil, A., Rodiac, R., Van Schaftingen, J.: Ginzburg–Landau relaxation for harmonic maps on planar domains into a general compact vacuum manifold. Archive for Rational Mechanics and Analysis, (Aug 2021)

  27. Monteil, A., Rodiac, R., Van Schaftingen, J.: Renormalised energies and renormalisable singular harmonic maps into a compact manifold on planar domains. Mathematische Annalen, May (2021)

  28. Nguyen, L., Zarnescu, A.: Refined approximation for minimizers of a Landau-de Gennes energy functional. Calc. Var. Partial Differential Equations 47(1–2), 383–432 (2013)

    Article  MathSciNet  Google Scholar 

  29. Sharpe, R.W.: Differential Geometry. Graduate Texts in Mathematics. Springer Verlag, (1997)

Download references

Acknowledgements

The first author was supported in part by NSF grant DMS-2106551.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitry Golovaty.

Additional information

Communicated by J. M. Ball.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

In this appendix we provide a short summary of previous results that are relevant to the present manuscript. We start with the following theorem that was proved in [13].

Theorem 5.1

Let \(g : \partial \Omega \rightarrow {\mathcal {P}}\) be a non-contractible curve in \({\mathcal {P}}\) and suppose that \(u_\varepsilon \in W^{1,2}(\Omega ; F_1)\) is a minimizer of \(E_\varepsilon \) among functions \(u \in W^{1,2}(\Omega ; F_1)\) that satisfy the Dirichlet boundary condition \(u = g\) on \(\partial \Omega .\) First, the minimizers \(u_\varepsilon \) take values in the convex envelope of \({\mathcal {P}}\); in particular they are uniformly bounded in \(\varepsilon \). Second, there is a single point a in the interior of \(\Omega \) such that the \(u_\varepsilon \) converge strongly (along a subsequence) to \(u_0 \in W^{1,2}(\Omega \setminus B_R\{a\}; {{\mathcal {P}}})\) in \(W^{1,2}(\Omega \setminus B_R\{a\}; F_1)\) as \(\varepsilon \rightarrow 0\) for any fixed \(R >0\). Finally, for any open set \(U \subset \subset {\overline{\Omega }}\setminus \{a\}\), \(u_0\) minimizes \(\int _U \left| {\nabla v}\right| ^2\) among functions \(v \in W^{1,2}_{loc}(\Omega \setminus \{a\}; {{\mathcal {P}}})\) satisfying \(v = u_0\) on \(\partial U\).

To describe the structure of \(u_0,\) let \(M_a^3({\mathbb {R}}^{})\) be the set of antisymmetric \(3\times 3\) matrices and let \([A;B]=AB-BA\) denote the commutator of matrices A and B. It turns out that one can consider a vector field \(j(u_0)\) with matrix entries

$$\begin{aligned} j(u_0) = \left( \left[ u_0 , \frac{\partial u_0}{\partial x}\right] , \left[ u_0 , \frac{\partial u_0}{\partial y}\right] \right) , \end{aligned}$$

instead of \(u_0\) because \(u_0\) can always be recovered from \(j(u_0)\) (the reason for this reduces to the following standard fact: if \(A:[0,T]\rightarrow M_a^3({\mathbb {R}}^{})\), then the solution of the initial value problem

$$\begin{aligned} \gamma ' = [\gamma , A],\ \ \gamma (0) \in {\mathcal {P}}, \end{aligned}$$

takes values in \({\mathcal {P}}\)). In light of this observation, the following theorem [13] gives a rough description of the limiting map \(u_0\) described in Theorem 5.1.

Theorem 5.2

Let \(u_0\) be as in Theorem 5.1. There is a function

$$\begin{aligned} \psi _0 \in (W^{1,2}\cap L^\infty )(\Omega ; M_a^3({\mathbb {R}}^{})) \end{aligned}$$

and a constant anti-symmetric matrix \(\Lambda _0\) such that

$$\begin{aligned} j(u_0) = \frac{1}{2 r} ({\hat{\theta }} \Lambda _0) + \nabla ^\perp \psi _0 \end{aligned}$$

in \(\Omega .\) Here \(a \in \Omega \) is as defined in Theorem 5.1, r and \({\hat{\theta }}\) are the radial variable and the unit vector in an angular direction for polar coordinates centered at a respectively, and we interpret \(({\hat{\theta }} \Lambda _0)\) and \(\nabla ^\perp \psi _0\) as matrix-valued vector fields according to (5.2) and (5.1) respectively. Further, \(\psi _0\) satisfies

$$\begin{aligned} \Delta \psi _0 = 2\left[ \frac{\partial \psi _0}{\partial x}, \frac{\partial \psi _0}{\partial y}\right] + \frac{1}{\pi r} \left[ \nabla \psi _0 \cdot {\hat{\theta }} , \Lambda _0\right] , \end{aligned}$$

in \(\Omega \), where we interpret \(\nabla \psi _0 \cdot {\hat{\theta }}\) according to (5.3), subject to boundary conditions

$$\begin{aligned} -\nabla \psi _0 \cdot \nu = \left[ g, \frac{dg}{d \tau }\right] - \frac{{\hat{\theta }}\cdot \tau }{2\pi r} \Lambda _0, \end{aligned}$$

on \(\partial \Omega ,\) where \(\nu \) and \(\tau \) are the outward unit normal and unit tangent vector to \(\partial \Omega ,\) respectively. Finally, the function \( Z_{u_0}(x) := \frac{1}{2\pi r}(\Lambda _0 - u_0 \Lambda _0 - \Lambda _0 u_0) \in L^2(\Omega ; M_a^3({\mathbb {R}}^{})). \)

Notice that in the preceding theorem we deal with matrix-valued functions \(u:\Omega \rightarrow M^3({\mathbb {R}}^{})\) and matrix-valued vector fields \(F : \Omega \rightarrow (M^3({\mathbb {R}}^{}))^2\), \(F = (F_1,F_2)\). For a matrix-valued function u, the gradient and its perpendicular are given by the matrix-valued vector fields

$$\begin{aligned} \nabla u = \left( \frac{\partial u}{\partial x}, \frac{\partial u}{\partial y} \right) , \,\,\, \nabla ^\perp u = \left( \frac{\partial u}{\partial y}, -\frac{\partial u}{\partial x} \right) , \end{aligned}$$
(5.1)

respectively. For matrix-valued vector fields, the divergence and curl

$$\begin{aligned} \nabla \cdot F = \frac{\partial F_1}{\partial x} + \frac{\partial F_2}{\partial y}, \,\,\, \nabla ^\perp \cdot F = \frac{\partial F_2}{\partial x} - \frac{\partial F_1}{\partial y} \end{aligned}$$

are matrix-valued functions. When \(z :\Omega \rightarrow {\mathbb {R}}^{2}\) and \(A:\Omega \rightarrow M^3({\mathbb {R}}^{})\), the matrix-valued vector field zA has the entries

$$\begin{aligned} zA = (z_1A, z_2A). \end{aligned}$$
(5.2)

On the other hand, if F is a matrix-valued vector field and \(e=(e_1, e_2)\in {\mathbb {R}}^{2}\), we set

$$\begin{aligned} F\cdot e = e_1F_1+e_2F_2, \end{aligned}$$
(5.3)

which is a matrix-valued function. We emphasize the difference between \(F\cdot e\) and zA defined in (5.2).

In the next proposition we summarize the properties of the potential we use.

Proposition 5.3

For \(u \in M^3_{s, 1}({\mathbb {R}}^{})\), define

$$\begin{aligned} W_\beta (u) = \frac{1}{4}(1-\left| {u}\right| ^2)^2 - \beta \, {\mathrm{det}}(u). \end{aligned}$$
(5.4)

We have

$$\begin{aligned} W_\beta (u) = \frac{1}{4}(1-\left| {u}\right| ^2)^2 - \frac{\beta }{6}(1 - 3\left| {u}\right| ^2 + 2{\mathrm{tr}}(u^3)). \end{aligned}$$
(5.5)

Next, for \(u \in M_{s, 1}^3({\mathbb {R}}^{})\) such that \(\langle u, P\rangle \ge 0\) for all \(P \in {\mathcal {P}}\), we have

$$\begin{aligned} \frac{(1-\left| {u}\right| ^2)^2}{3} \le \left| {u-u^2}\right| ^2 \le \frac{(1-\left| {u}\right| ^2)^2}{2}. \end{aligned}$$
(5.6)

Lastly, we have

$$\begin{aligned} W_\beta (u) \ge \frac{(3-\beta )}{6}({\mathrm{dist}}(u, {\mathcal P}))^2. \end{aligned}$$
(5.7)

Proof

We start by recalling that Cayley-Hamilton theorem for matrices \(u\in M_{s, 1}^3({\mathbb {R}}^{})\) tells us

$$\begin{aligned} u^3 - u^2 + \frac{(1-\left| {u}\right| ^2)}{2}u - {\mathrm{det}}(u)I = 0. \end{aligned}$$

Applying this in (5.4) gives us (5.5).

Next, again from Cayley-Hamilton theorem, for \(u \in M_{s, 1}^3({\mathbb {R}}^{})\), we obtain

$$\begin{aligned} \left| {u-u^2}\right| ^2 + 2{\mathrm{det}}(u) = \frac{(1-\left| {u}\right| ^2)^2}{2}. \end{aligned}$$
(5.8)

In particular, if \(u \in M_{s, 1}^3({\mathbb {R}}^{})\) has \(\langle u, P\rangle \ge 0\) for all \(P \in {\mathcal {P}}\), then

$$\begin{aligned} \frac{(1-\left| {u}\right| ^2)^2}{3} \le \left| {u-u^2}\right| ^2 \le \frac{(1-\left| {u}\right| ^2)^2}{2}. \end{aligned}$$

This is (5.6)

Under these conditions, if \({\mathrm{dist}}(u, {{\mathcal {P}}}) \le \delta \le \frac{1}{4}\), it is not hard to check that

$$\begin{aligned} \frac{2}{3}{\mathrm{dist}}(u, {{\mathcal {P}}}) \le 2\frac{\left| {u-u^2}\right| }{1-\delta } \le \frac{\left| 1-{|u|}^2\right| }{1-\delta } \end{aligned}$$

Furthermore, again for \(u \in M_{s, 1}^3({\mathbb {R}}^{})\) such that \(\langle u, P\rangle \ge 0\) for all \(P \in {\mathcal {P}}\), a lengthy, but ultimately straight forward minimization shows that

$$\begin{aligned} (1-\left| {u}\right| ^2)^2 \ge 12 \,\, {\mathrm{det}}(u). \end{aligned}$$

Hence, for \(1\le \beta < 3\), and \(u \in M_{s, 1}^3({\mathbb {R}}^{})\) such that \(\langle u, P\rangle \ge 0\) for all \(P \in {\mathcal {P}}\), we have

$$\begin{aligned} W_\beta (u)&= \frac{1}{4}(1-\left| {u}\right| ^2)^2 - \beta {\mathrm{det}}(u) = \frac{3-\beta }{12} ( 1 - \left| {u}\right| ^2)^2 + \beta ( \frac{1}{12}(1-\left| {u}\right| ^2)^2 - {\mathrm{det}}(u) )\\&\ge \frac{3-\beta }{12}(1-\left| {u}\right| ^2)^2 \ge \frac{(3-\beta )}{6}({\mathrm{dist}}(u, {{\mathcal {P}}}))^2. \end{aligned}$$

This is (5.7). \(\square \)

Remark 5.4

The potential \(W_{\beta }\) in [13] is written as

$$\begin{aligned} W_\beta (u) = \frac{1}{2}(1-\left| {u}\right| ^2)^2 - \beta \, {\mathrm{det}}(u). \end{aligned}$$

As can be seen from (5.8), in order for this potential to be equal \(\left| {u-u^2}\right| ^2\) we need to choose \(\beta =2\). In [13] we erroneously stated that our results were valid for \(2<\beta <6\) while the correct range should be \(2\le \beta < 6\). In this paper, however, we use the expression given in (5.4).

Proposition 5.5

Let \(\Omega \subset {\mathbb {R}}^{2}\) be a smooth, bounded, simply-connected open set, and \(u_\varepsilon : \Omega \rightarrow M_{s, 1}^3({\mathbb {R}}^{})\) be a minimizer of the LdG energy with non-contractible boundary data in \({\mathcal {P}}\). Let \(a\in \Omega \) be the distinguished point in \(\Omega \) that Theorem 5.1 shows exist. For \(r>0\) such that \(B_{2r}(a)\subset \Omega \), there is \(\varepsilon _0 > 0\) and a constant \(C>0\) such that

$$\begin{aligned} \left| {(\nabla u_\varepsilon )(x)}\right| + \frac{(1-\left| {u_\varepsilon }\right| ^2)}{\varepsilon ^2} \le C \end{aligned}$$

for all \(x\in \Omega \setminus B_r(a)\), and all \(0 < \varepsilon \le \varepsilon _0\).

Proof

The proof follows [3]. Let us observe that the end of the proof of Lemma 8 of [13] shows that minimizers \(u_\varepsilon \) satisfy

$$\begin{aligned} \limsup _{\varepsilon \rightarrow 0} \int _{\Omega \setminus B_r(a)} \frac{W(u_\varepsilon )}{\varepsilon ^2} = 0. \end{aligned}$$

We now appeal to Steps A.2 and B.2 of the proof of Theorem 1 of [2], to conclude that \(W(u_\varepsilon ) \rightarrow 0\) uniformly in \(\Omega \setminus B_r(a)\). In particular, for \(\delta > 0\) we can choose \(\varepsilon _0 > 0\) such that

$$\begin{aligned} 0 \le 1-\left| {u_\varepsilon ^2}\right| (x) \le \delta \end{aligned}$$

for all \(x\in \Omega \setminus B_r(a)\) and all \(0 < \varepsilon \le \varepsilon _0\).

We next recall from the appendix of [13] that

$$\begin{aligned} \frac{4-\beta }{\varepsilon ^2}(1-\left| {u}\right| ^2) - 4W_{\frac{3\beta }{4}} = -\Delta \frac{\left| {u}\right| ^2}{2} + \left| {\nabla u}\right| ^2 = -\langle u , \Delta u\rangle . \end{aligned}$$

We know from [13] that \(\langle u_\varepsilon , P \rangle \ge 0\) for all \(P \in {\mathcal {P}}\), so we deduce

$$\begin{aligned} 4W_{\frac{3\beta }{4}}(u_\varepsilon ) \le (1-\left| {u}\right| ^2)^2. \end{aligned}$$

Hence, we can choose \(\varepsilon _0 > 0\) small enough for

$$\begin{aligned} \frac{4-\beta }{\varepsilon ^2}(1-\left| {u_\varepsilon }\right| ^2) - 4W_{\frac{3\beta }{4}}(u_\varepsilon ) \ge \delta (1-\left| {u_\varepsilon }\right| ^2) \end{aligned}$$

in \(\Omega \setminus B_r(a)\), for all \(0 <\varepsilon \le \varepsilon _0\). Since \(\left| {u_\varepsilon }\right| \le 1\), we conclude that

$$\begin{aligned} \left| {\Delta u_\varepsilon }\right| \ge \delta (1-\left| {u_\varepsilon }\right| ^2) \end{aligned}$$

in \(\Omega \setminus B_r(a)\), for all \(0 <\varepsilon \le \varepsilon _0\).

From the Euler-Lagrange equation for \(u_\varepsilon \), we obtain

$$\begin{aligned} -\Delta \frac{\partial u_\varepsilon }{\partial x_k} + \frac{1}{\varepsilon ^2} (D^2_u W)(u)\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) = \frac{\partial \lambda _\varepsilon }{\partial x_k}I, \end{aligned}$$

and then

$$\begin{aligned} \Delta \left| {\nabla u_\varepsilon }\right| ^2 = 2\left| {D^2_x u}\right| ^2 + \frac{2}{\varepsilon ^2}\sum _{k=1}^2 \langle (D^2_u W)(u)\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle . \end{aligned}$$

Now, writing \(v_\varepsilon \) for the nearest element of \({\mathcal {P}}\) to \(u_\varepsilon \), we have

$$\begin{aligned} \langle (D^2_u W_\beta )(u_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle&= \langle (D^2_u W)(v_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle \\&\quad + \langle (D^2_u W_\beta )(u_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) - (D^2_u W_\beta )(v_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle . \end{aligned}$$

Now \(v_\varepsilon \in {\mathcal {P}}\), which is the set of minimizers of \(W_\beta \), and it is easy to check that \({\mathrm{tr}}(\frac{\partial u_\varepsilon }{\partial x_k}) = 0\). Hence

$$\begin{aligned} \langle (D^2_u W)(v_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle \ge 0. \end{aligned}$$

We deduce that

$$\begin{aligned} \langle (D^2_u W_\beta )(u_\varepsilon )\left( \frac{\partial u_\varepsilon }{\partial x_k}\right) , \frac{\partial u_\varepsilon }{\partial x_k} \rangle \ge -C\left| {u_\varepsilon - v_\varepsilon }\right| \left| {\frac{\partial u_\varepsilon }{\partial x_k}}\right| ^2 \ge -C(1-\left| {u_\varepsilon }\right| ^2)\left| {\frac{\partial u_\varepsilon }{\partial x_k}}\right| ^2, \end{aligned}$$

where the last inequality holds because from the comments before the proposition we have

$$\begin{aligned} \left| {u_\varepsilon -v_\varepsilon }\right| = {\mathrm{dist}}(u_\varepsilon , {{\mathcal {P}}}) \le C(1-\left| {u}\right| ^2). \end{aligned}$$

We conclude that

$$\begin{aligned} \Delta \left| {\nabla u_\varepsilon }\right| ^2 \ge 2\left| {D^2_x u}\right| ^2 - C\frac{(1-\left| {u_\varepsilon }\right| ^2)}{\varepsilon ^2}\left| {\nabla u}\right| ^2 \ge 2\left| {D^2_x u}\right| ^2 - C\left| {\Delta u_\varepsilon }\right| \left| {\nabla u}\right| ^2. \end{aligned}$$

Since this implies

$$\begin{aligned} \Delta \left| {\nabla u_\varepsilon }\right| ^2 \ge \left| {D^2_x u_\varepsilon }\right| ^2 - C\left| {\nabla u}\right| ^4 \end{aligned}$$

in \(\Omega \setminus B_r(a)\), we can apply Steps A.4 and B.3 of the proof of Theorem 1 of [2] to conclude that

$$\begin{aligned} \left| {\nabla u_\varepsilon }\right| \le C \end{aligned}$$

in \(\Omega \setminus B_r(a)\), for some constant independent of \(\varepsilon \in ]0, \varepsilon _0]\).

Finally, we recall from [13] that

$$\begin{aligned} \Delta \frac{\left| {u_\varepsilon }\right| ^2}{2} = \frac{4}{\varepsilon ^2} W_{\frac{3\beta }{4}}(u_\varepsilon ) -\frac{4-\beta }{\varepsilon ^2}(1-\left| {u_\varepsilon }\right| ^2) + \left| {\nabla u_\varepsilon }\right| ^2. \end{aligned}$$

From here, \(\zeta = 1-\left| {u_\varepsilon }\right| ^2\) satisfies

$$\begin{aligned} -\Delta \zeta + \frac{4-\beta }{\varepsilon ^2}\zeta = \left| {\nabla u_\varepsilon }\right| ^2. \end{aligned}$$

Steps A.5 and B.4 of Theorem 1 of [2] give us the last conclusion of the proposition. \(\square \)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Golovaty, D., Montero, J.A. Refined asymptotics for Landau-de Gennes minimizers on planar domains. Calc. Var. 61, 199 (2022). https://doi.org/10.1007/s00526-022-02306-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-022-02306-4

Mathematics Subject Classification

Navigation