Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Recent analytic development of the dynamic Q-tensor theory for nematic liquid crystals

  • Liquid crystals are a typical type of soft matter that are intermediate between conventional crystalline solids and isotropic fluids. The nematic phase is the simplest liquid crystal phase, and has been studied the most in the mathematical community. There are various continuum models to describe liquid crystals of nematic type, and Q-tensor theory is one among them. The aim of this paper is to give a brief review of recent PDE results regarding the Q-tensor theory in dynamic configurations.

    Citation: Xiang Xu. Recent analytic development of the dynamic Q-tensor theory for nematic liquid crystals[J]. Electronic Research Archive, 2022, 30(6): 2220-2246. doi: 10.3934/era.2022113

    Related Papers:

    [1] Yazhou Chen, Dehua Wang, Rongfang Zhang . On mathematical analysis of complex fluids in active hydrodynamics. Electronic Research Archive, 2021, 29(6): 3817-3832. doi: 10.3934/era.2021063
    [2] Haibo Cui, Junpei Gao, Lei Yao . Asymptotic behavior of the one-dimensional compressible micropolar fluid model. Electronic Research Archive, 2021, 29(2): 2063-2075. doi: 10.3934/era.2020105
    [3] Kai Jiang, Wei Si . High-order energy stable schemes of incommensurate phase-field crystal model. Electronic Research Archive, 2020, 28(2): 1077-1093. doi: 10.3934/era.2020059
    [4] Wenjuan Liu, Jialing Peng . Analyticity estimates for the 3D magnetohydrodynamic equations. Electronic Research Archive, 2024, 32(6): 3819-3842. doi: 10.3934/era.2024173
    [5] Zhefeng Xu, Xiaoying Liu, Luyao Chen . Hybrid mean value involving some two-term exponential sums and fourth Gauss sums. Electronic Research Archive, 2025, 33(3): 1510-1522. doi: 10.3934/era.2025071
    [6] Wen Teng, Xiansheng Dai . Nonabelian embedding tensors on 3-Lie algebras and 3-Leibniz-Lie algebras. Electronic Research Archive, 2025, 33(3): 1367-1383. doi: 10.3934/era.2025063
    [7] Xin Liu, Guang-Xin Huang . New error bounds for the tensor complementarity problem. Electronic Research Archive, 2022, 30(6): 2196-2204. doi: 10.3934/era.2022111
    [8] Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150
    [9] Deyue Zhang, Yukun Guo . Some recent developments in the unique determinations in phaseless inverse acoustic scattering theory. Electronic Research Archive, 2021, 29(2): 2149-2165. doi: 10.3934/era.2020110
    [10] Lot-Kei Chou, Siu-Long Lei . High dimensional Riesz space distributed-order advection-dispersion equations with ADI scheme in compression format. Electronic Research Archive, 2022, 30(4): 1463-1476. doi: 10.3934/era.2022077
  • Liquid crystals are a typical type of soft matter that are intermediate between conventional crystalline solids and isotropic fluids. The nematic phase is the simplest liquid crystal phase, and has been studied the most in the mathematical community. There are various continuum models to describe liquid crystals of nematic type, and Q-tensor theory is one among them. The aim of this paper is to give a brief review of recent PDE results regarding the Q-tensor theory in dynamic configurations.



    Liquid crystals are often considered as the fourth state of the matter besides the gas, liquid and solid, or as an intermediate state between liquid and solid. Liquid crystals are partially ordered materials that can translate freely as conventional fluid, while exhibit certain long-range order below a critical temperature. There are many different types of liquid crystals, the main classes being nematics, smectics and cholesterics. The nematic phase is the simplest among all liquid crystal phases. In this phase molecules float around as in a liquid phase, but have the tendency of aligning along a preferred direction due to their orientation [1], which makes the liquid anisotropic. The rod-like nematic liquid crystals are the most widely studied among all liquid crystals. The study of liquid crystals is of great significance for both fundamental scientific researches as well as their widespread applications in industry. Mathematical theories at different levels have been developed to describe nematic liquid crystals, ranging from the microscopic molecular theory to macroscopic continuum ones. Generally speaking, there are several closely related models in the continuum theory to describe nematic liquid crystals [2,3,4], namely Oseen-Frank theory [5], Ericksen's theory [3], and Landau-de Gennes theory [2,6]. In these theories, various order parameters are utilized to encode mathematically the local ordering of liquid crystal molecules.

    Assume liquid crystals are occupying an open region ΩRd, d=2,3. In the Oseen-Frank theory, the order parameter is

    n:ΩS2. (1.1)

    In the Ericksen's theory, the order parameter is

    (n,s)S2×[1/2,1]. (1.2)

    In the Landau-de Gennes theory, the order parameter is

    QQ:={MR3×3,Mt=M,tr(M)=0}, (1.3)

    which is an element in the space of Q-tensors. The Oseen-Frank theory can account for point defects but not line/surface defects [7] (roughly speaking, they are discontinuities in the alignment of liquid crystals). The Ericksen's theory incorporates both line/surface defects [8,9], but not biaxial liquid crystal structures. These two theories are vector theories in that the basic element is a unit vector n representing the mean orientation of neighbouring liquid crystal molecules. In the Landau-de Gennes theory, the basic element Q is a 3×3 symmetric, traceless matrix to describe the alignment of liquid crystal molecules. It is the most comprehensive one among these continuum theories [10,11].

    Concerning the mathematical analysis of nematic liquid crystals in the framework of these aforementioned continuum models, there has been a fast development particularly in recent years. The aim of this review is to provide an outline of the relevant PDE analysis on the dynamic Q-tensor theory, many of which are closely related to the author's own research work. Several intrinsically connected topics are presented in separate subsections below, and many unsolved problems are proposed. Due to spatial constraint, this review only aims to present major development of the analytic Q-tensor theory in dynamic settings, but not to provide a comprehensive list of the existing literature.

    The earliest PDE results on the dynamic Q-tensor theory came from [12,13], where the Cauchy problem of the hydrodynamic flow was studied. The corresponding coupled PDE system, called the Beris-Edwards system [14], consists of incompressible Navier-Stokes equations for the fluid velocity with highly nonlinear anisotropic force terms and nonlinear convection diffusion equations of parabolic type that describe the evolution of the Q-tensor:

    {tu+(u)u+P=νΔu+λ(τ+σ),u=0,tQ+uQS(u,Q)=ΓH(Q) (2.1)

    Here u(x,t):Rd×(0,+)Rd denotes the velocity field of the fluid, P:Rd×(0,+)R represents the hydrostatic pressure, and Q(x,t):Rd×(0,+)Q stands for the order parameter of liquid crystal molecules. The positive constants ν,λ,Γ stand for the fluid viscosity, the competition between kinetic energy and elastic potential energy, and macroscopic elastic relaxation time for the molecular orientation field, respectively. System (2.1) describes the interaction between the incompressible fluid flow and the alignment of liquid crystal molecules. The evolution of the fluid affects the direction of the molecules while changes in molecular alignment also influence the fluid velocity. The free energy of liquid crystal molecules in (2.1) is given by

    F(Q)=Rd[L2|Q|2+f(Q)]dx, (2.2)

    where the so called one constant approximation of the Oseen-Frank energy is used, and L>0 is an elastic constant. Meanwhile, the bulk part f(Q) is taken to be a fourth order degree polynomial:

    f(Q)=a2tr(Q2)b3tr(Q3)+c4tr2(Q2), (2.3)

    where a,b,c are assumed to be material and temperature dependent constants. Note that the assumption

    c>0 (2.4)

    is always imposed in all existing literature to ensure the free energy is bounded from below. The tensor H(Q) is defined to be the variational derivative of F(Q) with respect to Q under both symmetry and traceless constraints:

    H(Q)=F(Q)Q=LΔQaQ+b(Q213tr(Q2)I3)cQtr(Q2). (2.5)

    And the matrix-valued function S(u,Q) reads

    S(u,Q):=(ξD+ω)(Q+13I3)+(Q+13I3)(ξDω)2ξ(Q+13I3)tr(Qu). (2.6)

    Here

    D=u+Tu2,ω=uTu2

    represent the symmetric and skew-symmetric parts of the rate of strain tensor, respectively. In addition, S(u,Q) accounts for the rotating and stretching effects on the order parameter Q due to the fluid, as the liquid crystal molecules can be tumbled and aligned by the flow. The constant parameter ξR depends on the molecular shapes of the liquid crystal, which also measures the ratio between the tumbling and the aligning effects that a shear flow exerts on the liquid crystal director. In particular, when ξ=0 the system (2.1) is called the co-rotational Beris-Edwards system. While when ξ0, it is at times referred to as the non co-rotational Beris-Edwards system, or the full Beris-Edwards system.

    The symmetric and skew-symmetric parts of the stress terms in the first equation of (2.1) caused by anisotropy of liquid crystal molecules are given respectively by

    τ:=ξ(Q+13I3)H(Q)ξH(Q)(Q+13I3)+2ξ(Q+13I3)tr(QH(Q))LQQ, (2.7)
    σ:=QH(Q)H(Q)Q. (2.8)

    It is noted that the Beris-Edwards system (2.1) satisfies a dissipative energy law under various suitable boundary conditions:

    ddtE(u,Q)=ν2|u|dxλΓ|H(Q)|2dx. (2.9)

    Here E(u,Q) denotes the total energy of the hydrodynamic system, which is the sum of the Q-tensor free energy F and the kinetic energy 1/2|u|2dx.

    The Beris-Edwards system (2.1) contains a significant number of highly nonlinear terms and generates considerable analytical difficulties, mainly due to nonlinear coupling with the incompressible Navier-Stokes equations. Since it contains the incompressible Navier-Stokes equations as a subsystem, we may not expect any better well-posedness results than the Navier-Stokes equations. For the co-rotational case ξ=0, the first contribution came from [13], where the authors proved the existence of global weak solutions to the Cauchy problem in Rd for d=2,3.

    Definition 2.1. A pair (u,Q) is called a weak solution of the problem (2.1) in Rd, subject to initial data

    u(0,x)=u0(x)L2(Rd),u0=0inD(Rd),Q(0,x)=Q0(x)H1(Rd) (2.10)

    provided

    uLloc(0,+;L2)L2loc(0,+;H1),QLloc(0,+;H1)L2loc(0,+;H2),

    and for compactly supported test functions ψC([0,)×Rd;Rd), ψ=0, ϕC([0,)×Rd;Q), it holds

    0Rd(utψuu:u+νu:ψ)dxdtRdu0(x)ψ(0,x)dx=L0Rd[(QQ:ψ)(QΔQΔQQ):ψ]dxdt,0Rd[Q:tϕΓLΔQ:ϕQ:(uϕ)ωQ:ϕ+Qω:ϕ]dxdt=Γ0Rd[aQ+bQ2btr(Q2)dIdcQtr(Q2)]:ϕdxdt+RdQ0(x):ϕ(0,x)dx.

    Here A:B stands for the inner product between two matrices of the same size.

    Theorem 2.1. For the co-rotational Beris-Edwards system (2.1), there exists a global weak solution subject to the initial condition (2.10).

    The proof of Theorem 2.1 was composed of several steps. First, a regularized sequence (u(n),Q(n)) was constructed to an approximate system following the classical Friedrich's scheme. Next, uniform apriori bounds collected from the basic energy dissipative law as

    supnu(n)L(0,T;L2)L2(0,T;H1)<,supnQ(n)L(0,T;H1)L2(0,T;H2)<,

    combined with classical Aubin-Lions compactnes arguments, paved the way to pass limit n+.

    Meanwhile, the authors in [13] also obtained results on higher global regularity of solutions as well as the weak-strong uniqueness for d=2:

    Theorem 2.2. For any s>1 and initial data (u0,Q0)Hs(R2)Hs+1(R2), the co-rotational Beris-Edwards system admits a global strong solution in the sense that

    uL2loc(0,+;Hs+1(R2))Lloc(0,+;Hs(R2)),QL2loc(0,+;Hs+2(R2))Lloc(0,+;Hs+1(R2)).

    The proof of Theorem 2.2 was mainly based on Hs energy estimates, and the most crucial step to control the highest order nonlinear term was the discovery of the following cancellation relation

    Rdtr((ωQQω)ΔQ)dxRd(QΔQΔQQ)udx=0,

    for symmetric, smooth matrix-valued functions Q,Q:RdRd×d. The Littlewood-Paley approach was used because it allows for initial data (u0,Q0)Hs(R2)Hs+1(R2) for any s>1.

    Shortly afterwards, the uniqueness result in [13] was improved in [15] where the author proved uniqueness of global weak solutions for d=2. Besides, an alternative approach that combined the Friedrichs scheme and the Schaefer's fixed point theorem was used in [15] to prove the existence of global weak solutions to (2.1). Next, by using a Fourier-splitting method, the asymptotic behavior of the weak solution to the Cauchy problem in R3 was given in [16], with some extra assumptions on the initial data and coefficients in (2.3). Further, a partial regularity result of suitable weak solutions in R3 was achieved in [17]:

    Theorem 2.3. For initial data u0¯{vC0(R3),v=0}L2 and Q0H1(R3), there exists a global suitable weak solution (u,Q) of the co-rotational Beris-Edwards system (2.1). Further,

    (u,Q)C((0,+)×R3Σ),

    where Σ(0,+)×R3 is a closed set whose 1-dimensional Hausdorff measure is 0.

    The existence of weak suitable solutions can be constructed following the classical construction of its counterpart in the incompressible Navier-Stokes equations. And the proof of the partial regularity result in Theorem 2.3 involved several crucial steps. The first key point was to establish a weak maximum principle of Q for the suitable weak solution (u,Q), see also [18]:

    Lemma 2.1. Let (u,Q)L2((0,+);H1)×L2((0,+);H2) be a weak solution of the co-rotational Beris-Edwards system (2.1). If Q0L(R3), then there exists a constant C>0 that depends on Q0L(R3) and a,b,c only, such that

    |Q(t,x)|C,(t,x)(0,+)×R3.

    The second key point was to use a blowing up argument to obtain an ϵ0-regularity criterion of any suitable weak solution based on a local energy inequality. It states that if the suitable weak solution satisfies

    r2Pr(t0,x0)(|u|3+|Q|3)dtdx+(r2Pr(t0,x0)|P|32dtdx)2ϵ30,

    then (t0,x0)(0,)×R3 is a smooth point of (u,Q). The third key point was to achieve the higher order bound of (u,Q) near (t0,x0), which can be done by intrinsic cancellation property of several coupling nonlinear terms.

    On the other hand, the corresponding initial boundary value problems of the co-rotational Beris-Edwards system subject to various boundary conditions for d=2,3 were studied in [18,19,20], where the existence of global weak solutions, existence and uniqueness of local strong solutions as well as some regularity criteria were provided.

    Concerning the full Beris-Edwards system (2.1) with general ξR, its mathematical analysis is more difficult because when ξ0, there are higher order derivatives in (2.7) as well as (2.6) compared to the case when ξ=0. In [12], the authors proved the existence of global weak solutions for the Cauchy problem in Rd with d=2,3 for sufficiently small |ξ|. Note that such smallness assumption was necessary for the Cauchy problem otherwise in an infinite domain the Lp a priori estimates cannot be established. The authors also established the existence of global strong solutions for d=2 as that in Theorem 2.2. In their proof a differential inequality relating higher order norms of the solution was rather untraditional in the sense that it was with a double-logarithmic correction. To retrieve this higher order energy differential inequality for d=2, the following logarithmic embedding of H1+ϵ in L in conjunction with the precise growth of Sobolev embedding constant of H1 in any Lp played an essential role:

    QL(R2)QH1(R2)ln(e+Q2Hs(R2)QH1(R2)), (2.11)
    gL2p(R2)C(p)pg1pL2(R2)g11pL2(R2). (2.12)

    The uniqueness of weak solutions for d=2 was given in [21]. Later on, the existence and long time behavior of global strong solutions to (2.1) with any arbitrary ξR under periodic settings for d=2 were proved in [22]. Further, existence of global weak solutions and local well-posedness with higher time-regularity for the initial boundary value problem subject to inhomogeneous mixed Dirichlet/Neumann boundary conditions were given in [23]. Besides, a rigorous derivation from the Beris-Edwards system to the classical Ericksen-Leslie system [24,25] was provided in [26] by using the Hilbert expansion method. We want to point out that there has been a large number of literature on the analytic study of the simplified/full Ericksen-Leslie system, and interested reader may see [27] and references therein for more details. One unsolved regularity issue for the Beris-Edwards system is

    Unsolved research problem 1: Determine the size of the singular set of a suitable weak solution to the non co-rotational Beris-Edwards system (2.1) for any ξ0.

    Any suitable weak solution to (2.1) is a weak solution that additionally satisfies a local energy inequality [17], which turns out to be a necessary condition for smoothness of the solution. Compared to the proof in [17] for the co-rotational Beris-Edwards system, one essential difficulty in this research problem is that the weak maximum principle is not valid for Q. Specifically, the L norm of Q is not ensured to be bounded in terms of that of initial data Q0 as time evolves for any ξ0. This also has strong physical implications in that when ξ=0, the fluid flow only has tumbling but no stretching effect on liquid crystal molecules, while when ξ0 both effects would be imposed. A closely related but more subtle issue is the preservation of the initial eigenvalue range [28]. It is essentially the behaviour of eigenvalues of the Q-tensors under the fluid dynamics as time evolves.

    Such eigenvalue preservation property was proved in [29] for the co-rotational system under periodic settings:

    Theorem 2.4. Let T>0, ξ=0, aR, b>0, c>0 and (u0,Q0)H1(T2)×H2(T2) with u=0. Suppose the eigenvalues of the initial datum Q0 satisfy

    λi(Q0(x))[b+b24ac12c,b+b24ac6c],1i3,xT2 (2.13)

    and

    |a|<b23c.

    Denote (u,Q) the unique global strong solution to the Beris-Edwards system (2.1) on [0,T] with initial data (u0,Q0). Then the eigenvalues of Q(t,x) stays in the same interval as in (2.13) for any t[0,T] and xT2.

    The result in Theorem 2.4 is also valid in the whole space or for local strong solutions for d=3. The main idea to prove the preservation of the convex hull of eigenvalues of Q0 for any regular enough solution was based on an operator splitting and a nonlinear Trotter product formula. To be more precise, consider S(t,ˉS), or S(t,)ˉSQ the flow generated by the ODE system

    {ddtS=aS+b[S213tr(S2)I3]ctr(S2)S,S(t,)=ˉSQ. (2.14)

    And for ˉRH2(T2Q) denote R(t;s,ˉR)=V(t,s)ˉR the unique solution to the linear non-autonomous problem:

    {tRεΔR=uR+ωRRω,R(s,)=ˉR. (2.15)

    Then it aims to show both the solution S(t,ˉS) of the ODE system (2.14), and the two parameter evolution system V(t,s) of the problem (2.15) preserve the above closed convex hull of the range for the initial data in (2.13). And eventually, a nonlinear Trotter product formula provides a way of expressing the solution Q(t,x) of (2.1) as a limit of successive superpositions of solutions to the ODE system (2.14) and the problem (2.15).

    If the velocity u is neglected, then the Q-equation in (2.1) is reduced to a gradient flow generated by the free energy (2.2). As a consequence, the gradient flow can be "splitted" into a heat flow and an ODE system so that the initial convex hull of eigenvalues are proved to be preserved by both sub-flows, then the combination is performed by a Trotter product formula [28]. But the structure of the Q-equation in (2.1) is much more complicated due to the existence of fluid velocity. To overcome the difficulty, a Maier-Saupe type singular potential was introduced in [29] (see the later subsection for more details on it), whose special analytic properties were exploited. An alternative proof was given in [30] on a smooth, bounded domain with an extra assumption a0.

    On the other hand, it was shown in [29] the eigenvalue-range preservation of initial data Q0 cannot be always valid for strong solutions of the non co-rotational system for all ξ0. The main idea of the proof was to use contradiction argument and consider the so-called high Ericksen number limit, in which the limiting case is a weakly coupled system consisting an incompressible Euler equation for the fluid velocity with a reaction-convection equation for Q. It remains to study

    Unsolved research problem 2: Determine whether the eigenvalue-range preservation property is not true for the non co-rotational Beris-Edwards system for any ξ0.

    As pointed in [29], we expect in general for any ξ0 one does not have the preservation of initial eigenvalue-range. In fact, for a shear flow in the whole space, the coupled system simply reduces to an ODE system. This situation was analyzed in [31] and the results obtained there allow to conclude that the eigenvalue-range preservation does not hold in general circumstances. But this shear flow contains nonphysical aspects in that it is with infinite energy. Therefore, to address this problem, a crucial step is to construct a family of exact solutions to the system with finite energy.

    Other than the Beris-Edwards system (2.1), the inertial Qian-Shen model [32] is another hydrodynamic system that has recently aroused the interest of the PDE community. It reads

    {tu+uu+P=β42+(LQQ+β1Qtr(QD)+β5AD+β6QD)+(μ22(˙Q[ω,Q])+μ1[Q,(˙Q[ω,Q])]),u=0,J¨Q+μ1˙Q=LΔQaQ+b(Q21d|Q|2Id)cQ|Q|2˜μ22D+μ1[ω,Q]. (2.16)

    Here in (2.16),

    ˙Q=(t+u)Q

    denotes the material derivative the Q-tensor and for any two d×d matrices A,B, [A,B]:=ABBA denotes their commutator. Here J>0 in the Q-equation of (2.16) represents the inertial density. It is noted that this inertial term plays a conceivable role when the anisotropic axis is subject to large accelerations, as motivated by the director model [33].

    The hydrodynamic system (2.16) can be considered as the Q-tensor version of the full Ericksen-Leslie model (with an inertial term). A rigorous justification from the inertial Qian-Shen model (2.16) to the full Ericksen-Leslis system with an inertial term was performed in [34]. Compared to (2.1), the most specific feature of this model is the presence of the inertial term J¨Q that appears as a second-order material derivative in the Q-equation. This very term provides a hyperbolic character to the system of equations and is the main source of challenges in its PDE analysis. As an initial attempt to tackle with the system (2.16), a basic energy dissipative law to the inertial Qian-Shen model (2.16) was found in [35]:

    Theorem 2.5. Let d=2* or 3, and the coefficients in the system (2.1) satisfy

    *In [35] for d=2, Q is assumed to a tensor-valued function into the set of 2-dimensional Q-tensors, that is, symmetric and tracless 2×2 matrices
          {MR2×2|Mt=M,tr(M)=0}.

    β10,β40,μ10,β6β5=μ2,β5+β6=0,˜μ2=μ2.

    Then there exists a constant Cd that depends on μ2,β5,β6,μ2, such that for any classical solutions that decay fast enough at infinity it holds

    ddtRd[12|u|2+J2|˙Q|2+L2|Q|2+f(Q)]dx0, (2.17)

    provided β4>Cd. Further, for any T>0 the following a priori bounds can be established

    uL(0,T;L2(Rd))L2(0,T;H1(Rd))QL(0,T;H1(Rd))with˙QL(0,T;L2(Rd)),

    provided that

    1. d=2, a0.

    2. or there exist ˉμ1=ˉμ1(a,b,c), J0=J0(ˉμ1,a,b,c) and ˜Cd=˜Cd(μ2,β5,β6)>0 such that μ1>ˉμ1, J<J0, β4>˜Cd.

    However, one cannot construct weak solutions by virtue of these a priori estimates in Theorem 2.5. The most common approach for the construction of weak solutions is to use the compactness method, that is, to construct approximate sequence of solutions with same energy bounds and then pass to the limit. The main difficulty lies in the stress term QQ in that when J>0 the Q-equation is of hyperbolic nature, and it does not allow to collect enough a priori bounds, especially the L2(0,T;H2(Rd)) for Q. Alternatively, the global-wellposedness of the Cauchy problem to (2.16) for small initial data was established in [35]:

    Theorem 2.6. Let J<J0, μ1>ˉμ1, β4>˜Cd, where J0,ˉμ1,˜Cd are explicitly computable coefficients. Assume β1>0, μ1>0, a>0, and (u0,Q0)Hs(Rd)×Hs+1(Rd) with s>d2 for d=2,3. Then there exists ε0>0, depending on s and d such that if

    η0:=u0Hs+Q0Hs+1+˙Q0Hs<ε0,

    then there exists a unique global strong solution (u,Q) of (2.16). Furthermore, there exists C>0 that is independent of the solution such taht

    uL(0,;Hs(Rd))+uL2(0,;Hs(Rd))+QL(0,;Hs+1(Rd))+QL2(0,;Hs+1(Rd))˙QL(0,;Hs(Rd))+˙QL2(0,;Hs(Rd))Cε0.

    The major difficulty of the proof comes from the second order material derivative ¨Q that reads after expansion

    ¨Q=2tQ+2utQ+tuQ+(uu)Q+u2Qu,

    where tu and u2Qu are competing terms with regularizing term LΔQ in the Q-equation. To overcome the analytic difficulty, the major strategy was to stay as close as possible to standard cancellation with respect to convective derivatives. Furthermore, a higher order commutator estimate at the level of homogeneous Sobolev spaces ˙Hs was considered. To prove the existence of classical solutions, the authors managed to achieve a uniform energy estimate for approximate solutions in the form of

    Φ(t)+Ψ(t)CΦ(t)Ψ(t),

    where C>0, Φ controls the Hs-norms for the solution and Ψ is integrable function (in time) that involves Hs-norms. Then smallness assumption on Φ(0) paved the way to uniform bounds for Hs-norms of approximate solutions. The subsequent proof was completed by a compactness argument.

    The authors in [35] also constructed one example of the so-called twist wave solution that is a classical solution of the coupled system for which the flow velocity is identically zero [36]. The twist wave solution is a solution of (2.16) with zero flow, such that u=ω=D=0, ˙Q=tQ, but it satisfies a nonlinear constraint

    P=(QQ+μ22tQ+μ1[Q,tQ]).

    Afterwards, a finer analysis of coefficient assumptions used in entropy inequality and energy dissipation in Qian-Shen model was performed in [37]. To be more specific, for any integer s>d2+1 (d=2 or 3), denote

    Ein:=u02Hs+J˙Q02Hs+LQ02Hs+aQ02Hs,

    then a large initial data/local well-posedness result was established:

    Theorem 2.7. Suppose the coefficients of (2.16) satisfy

    {a>0,β1>0,β4>0,μ1>0,β5+β6=0. (2.18)

    If β6β5=μ2, the initial energy Ein<, and the so called Condition (H) is valid:

    (˜μ2μ2)2+4μ22<8β4μ1, (2.19)

    then the Cauchy problem of (2.16) admits a unique local strong solution (u,Q) that satisfies

    {uL(0,T;Hs(Rd))L2(0,T;˙Hs+1(Rd)),˙QL(0,T;Hs(Rd))L2(0,T;Hs(Rd)),QL(0,T;Hs+1(Rd)). (2.20)

    And the energy bound depends only on the initial data, the coefficients of the system, and T.

    At the same time, the authors in [37] also showed that when the relation β5+β6=0 and Condition (H) are not simultaneously satisfied, then smallness assumption must be imposed on initial data to achieve even local well-posedness following energy methods.

    Meanwhile, the global existence of dissipative solutions to a dissipative version of Qian-Shen model was established in [38]. Besides, we refer interested readers to [39], where the authors made a comprehensive study of the full Ericksen-Leslie system with an inertial term for d=1. It remains to solve the following two fundamental analytic issues to the inertial Qian-Shen model (2.16)

    Unsolved research problem 3: Prove the existence of global weak solutions for the inertial Qian-Shen model.

    So far there has been no systematic approach to handle such a hydrodynamic system with hyperbolic structures. New ideas or tools would be necessary.

    Unsolved research problem 4: Determine whether singularity exists for solutions of the inertial Qian-Shen model for large enough initial data.

    The counterpart of the inertial Qian-Shen model in the director theory is the full Ericksen-Leslie system with an inertial term [33]. In the Poiseuille flow of the full Ericksen-Leslie system with an inertial term, the solution takes the form

    u(x,t)=(0,0,u(x,t))t,n(x,t)=(sinθ(x,t),0,cosθ(x,t))t,

    where u and θ are scalar functions. Consequently, the full Ericksen-Leslie system is reduced to

    {ρut=˜a+(g(θ)ux+h(θ)θt)x,νθt+γ1θt=c(θ)(c(θ)θx)xh(θ)ux, (2.21)

    where ˜aR is the gradient of pressure along the flow direction, and

    {g(θ):=α1sin2θcos2θ+α5α22sin2θ+α3+α62cos2θ+α42,c(θ):=K1cos2θ+K3sin2θ,K1K3,h(θ):=γ1+γ2cos(2θ)2.

    However, such singularity formation proof cannot be applied to (2.16) since during the reduction of Q-tensor elastic energy to the Oseen-Frank energy K1K3 even if more quadratic terms of gradient of Q are added (see the discussion after (2.35) for details).

    Concerning the free energy (2.2), an interesting topic is the rigorous study of nematic-isotropic phase transition based on a natural Q-tensor gradient flow generated by it. Phase transitions between different phases of liquid crystals give rise to a variety of mathematical questions of great interest. Precisely speaking, for the free energy

    Fε(Q)=Ω[ε2|Q|2+1εf(Q)]dx, (2.22)

    where ΩRd is a smooth, bounded domain, ε>0 denotes the relative intensity of elastic and bulk energy, it is of importance to investigate the small-ε limit of its natural gradient flow dynamics with initial data undergoing a sharp interfacial transition:

    {tQε=ΔQε1ε2(aQεbQ2ε+b3|Qε|2I3+c|Qε|2Qε),inΩ×(0,T),Qε(x,0)=Q0ε(x),inΩ,Qε(x,t)=0,onΩ×(0,T). (2.23)

    It is known that all critical points of f(Q) are uniaxial [65] in the sense that

    Q=s(nn13I3),for somenS2,sR. (2.24)

    Besides, f(Q) has two families of stable local minimizers corresponding to the following choices of s=s±:

    s=0,s+=b+b224ac4c. (2.25)

    And in [40,41] the following bistable case is considered

    b2=27ac,a>0,c>0, (2.26)

    which from the physics point of view corresponds to the critical temperature at which the system favors the nematic and isotropic phases equally [2]. In this case, the two families of minimizers associated with (2.25) are the only global minimizers of f(Q) in the sense that

    f(Q)0,and the equality holds if and only ifQ{0}N,

    with

    N:={QQ|Q=s+(nn13I3),nS2}. (2.27)

    In [40,41], different approaches were utilized to such rigorous proof under the assumption that the initial data of the gradient flow undergoes a sharp transition near a smooth surface. To be more specific, let T>0,

    I=t[0,T](It×{t})is a smoothly evolving closed surface inΩ, (2.28)

    which starts from a closed smooth surface I0Ω. Denote Ω+(t) the domain enclosed by It, and d(x,It) the signed distance from x to It that takes positive values in Ω+(t) and negative values in Ω(t)=Ω¯Ω+(t), i.e.,

    Ω+(t):={xΩ|d(x,It)>0},Ω(t):={xΩ|d(x,It)<0}. (2.29)

    At the same time, we define the distorted parabolic cylinder by

    Ω±T:=t(0,T)Ω±(t)×{t}. (2.30)

    Moreover, for any sufficiently small δ>0, we denote the δ-neighborhood of It by

    It(δ):={xΩ:|d(x,It)|<δ}.

    As a consequence, there exists a suitably small constant δI(0,1) such that the nearest point projection

    PI(,t):It(δI)It

    is smooth for any t[0,T], while the interface I stays away the boundary Ω for at least δI.

    To introduce the modulated energy Fε[Qε|I] for the problem (2.23), let us extend the inner normal vector field nI of It to a neighborhood by

    ξ(x,t):=η(d(x,It))n(PI(x,t),t),

    where ηCc(R) is a cutoff function with the following properties:

    {η(z)=η(z),zR,η(z)0,z[0,+),η(z)=1z2,for |z|δI2,η(x)=0,for |z|δI.

    Further, dF is defined to be the quasi-distance function

    dF(Q):=inf{102F(γ(t))|γ(t)|dt|γC0,1([0,1],Q),γ(0)N,γ(1)=Q}

    Consequently, the modulated energy is Fε is defined to be

    Fε[Qε|I](t):=Ω[ε2|Qε(,t)|2+1εF(Qε(,t))ξψε(,t)]dt, (2.31)

    with

    ψε(x,t):=(ϕεdF)Qε(x,t),

    and the convolution is considered to be in the Q-tensor space Q with a standard mollifier ϕε.

    The main result in [41] states that

    Theorem 2.8. Suppose the surface It evolves by mean curvature flow, and it encloses a simply connected domain Ω+(t) for any t[0,T]. Let the initial data Qinε of (2.23) be well-prepared such that

    εQ0εL(Ω)+Fε[Qε|I](0)C0ε,

    where Fε[Qε|I] is the modulated energy of Fε given in (2.31), and C0>0 does not depend on ε. Then it holds up to a subsequence that

    Qεk˜Q:=s±(n(x,t)n(x,t)13I3)strongly inC([0,T];L2loc(Ω±(t)))

    where s± are given in (2.25). Furthermore, n(x,t)H1(Ω+T;S2) is a harmonic map heat flow onto S2 with homogeneous Neumann boundary conditions.

    Theorem 2.8 mainly states that starting from Q0ε with a nematic-isotropic transition from the nematic regime Ω+(0) into the isotropic regime Ω(0), the solution Qε of the gradient flow (2.23) will converge to ˜QN that takes values in nematic phase in Ω+(t), and to the isotropic phase 0 in Ω(t). The interface between Ω+(t) and Ω(t) evolves by mean curvature flow and the limit map ˜Q in Ω+(t) is a harmonic map heat flow into N. To prove Theorem 2.8, approaches involving matched asymptotic expansions and spectral gap estimates were used in [40] (with slightly different setting of initial data other than that in Theorem 2.8), while the methods of weak convergence together with modulated energy method were adopted in [41].

    However, this topic becomes more challenging if more anisotropic elastic terms are included in the free energy:

    [L1|Q|2+L2jQikkQij+L3jQijkQik+f(Q)]dx. (2.32)

    Here and after, L1,L2,L3 are material-dependent elastic constants, kQij stands for the k-th spatial partial derivative of Qij, and Einstein summation convention is used.

    Unsolved research problem 5: Under proper assumptions on L1,L2,L3, study the scaling limit of the L2 gradient flow generated by the free energy (2.32) instead of F.

    For the co-rotational Beris-Edwards system with three elastic constants L1,L2,L3 (in which (2.2) is replaced by (2.32)), the Cauchy problem with d=3 was studied in [54], where the authors proved the existence of global weak solutions as well as the existence and uniqueness of global strong solutions provided that the fluid viscosity is sufficiently large. For the full Beris-Edwards system that involves three elastic constants L1,L2,L3 with general ξR, the corresponding initial boundary value problem was studied in [61], where the local well-posedness result was significantly improved. It states that

    Theorem 2.9. For any smooth bounded domain ΩR3, consider the full Beris-Edwards system with three elastic constants L1,L2,L3 subject to the initial condition

    (u0,Q0)H10(Ω)×H2(Ω),u0=0

    and the boundary condition

    (u,Q)|Ω=(0,Q0|Ω).

    suppose the elastic constants satisfy

    L1>0,L1+L2+L3>0,

    and the initial data (u0,Q0) satisfies certain compatibility conditions. Then for some T>0, the system admits a unique strong solution in the sense that

    uH2(0,T;H1(Ω))H1(0,T;H1(Ω))L(0,T;H2(Ω)),QH2(0,T;L2(Ω))H1(0,T;H2(Ω))L(0,T;H3(Ω)).

    Moreover, a strong Legendre condition was discovered in the Euler-Lagrange equation associated with the free energy (2.32). Theorem 2.9 is significant in that it greatly improves the spatial regularity of solutions established in [23], and it extends the result in [23] to the anisotropic elastic energy case that is far more involved than the isotropic elastic energy only. The proof of Theorem 2.9 replies on the observation that terms containing third derivatives of Q in stress tensors can be eliminated, and the system can be transformed into a Stokes-type system with positive definite viscosity coefficients.

    We proceed to be focused on the free energy with a cubic L4 term [6,10,11]:

    ˜F(Q)=[L1|Q|2+L2jQikkQij+L3jQijkQik+L4QlkkQijlQij+f(Q)]dx (2.33)

    The appearance of this cubic L4 term is motivated by the fact that it allows a complete reduction of the elastic energy part in (2.33) to the classical Oseen-Frank energy, by formally taking [28,46]

    Q(x)=s+[n(x)n(x)13I3],s+R. (2.34)

    To be more specific, inserting (2.34) into (2.33), we recover the Oseen-Frank energy (with four elastic constants) [5]

    {K1(divn)2+K2(ncurln)2+K3|n×curln|2+(K2+K4)[tr(n)2(divn)2]}dx, (2.35)

    where [10]

    {K1=2L1s2+L2s2+L3s223L4s3,K2=2L1s223L4s3,K3=2L1s2+L2s2+L3s2+43L4s3,K4=L3s2.

    If the L4 cubic term is not present, say L4=0, then this reduction is incomplete because K1K3 while K1,,K4 are independent constants. However, the retention of this term makes the free energy ˜F unbounded from below [44].

    The physical relevance of the free energy (2.33) under 2D settings was investigated in [28]. That is, Q is assumed to be a tensor-valued function into the space of 2-dimensional Q-tensors

    {MR2×2|Mt=M,tr(M)=0}. (2.36)

    In particular, an L2 gradient flow for d=2 into the space (2.36) associated with (2.33) was proposed:

    Qt=δ˜FδQ+λI2+μμt, (2.37)

    where λR is a Lagrange multiplier corresponding to the constraint tr(Q)=0 and μR2×2 is a Lagrange multiplier associated with the constraint Qt=Q. After expansion it yields

    Qijt=2L1ΔQijaQijctr(Q2)Qij+(L2+L3)(jkQik+ikQjk)(L2+L3)lkQlkδij+2L4lQijkQlk+2L4lkQijQlkL4iQkljQkl+L42|Q|2δij. (2.38)

    The problem (2.38) is studied on a bounded domain ΩR2 with the following initial data and boundary conditions:

    Q(x,0)=Q0(x),andQ(x,t)|Ω=˜Q(x). (2.39)
    Q0|Ω=˜Q.

    The main aim of [28] was two-folded: show the existence of global weak solutions to the problem (2.38)-(2.39) with H1L initial data that is small in L and finite time blow up (in L2) of solutions for specially constructed large initial data. To prove the global existence, the main difficulty comes from the fact that the energy is a priori unbounded from below. But it follows directly from (2.33) that if QL is suitably small, then the cubic L4 term can be absorbed into the other quadratic terms which are positive definite under mild coercivity assumptions. In this way the H1-level energy in the gradient flow can be effectively used, provided one can a priori ensure a smallness condition on the L-norm. To this end, a key lemma was established:

    Lemma 2.2. For the 2D evolution problem (2.38)-(2.39) on a bounded smooth domain ΩR2, suppose

    L1+L2>0,L1+L3>0. (2.40)

    For any smooth solution Q there exists an explicitly computable constant η1 (depending on Li,i=1,,4) such that if

    ˜QL(Ω)Q0L(Ω)<2η1

    and

    |a|2cη1,

    then for any T>0, we have

    QL((0,T)×Ω)2η1.

    Lemma 2.2 plays an essential role in the proof of existence of global weak solutions in that it ensures the smallness of QL globally in time, provided it is suitably small at initial time. Together with the energy dissipative law inherited from the gradient flow, one can establish the uniform a priori bounds for QL(0,T;H1) and QL2(0,T;H2). But the highly nonlinear terms related to L4 makes the approximation scheme to establish global weak solutions nonstandard, even if all the necessary a priori bounds are collected. In particular, in order to obtain coercivity of the second order terms, QL must be kept small in the approximation scheme. This is completely nontrivial since Q is a 2×2 matrix instead of a scalar function. To achieve this goal, the following singular potential was introduced:

    ˜f(Q):={ln(8η2|Q|2),if |Q|2<8η2,+,otherwise.

    for some suitably small constant η2>0. Correspondingly, an extra term

    ε˜fQ,

    was added to (2.38), which enforced the approximate system to admit |Q|2<8η2 almost everywhere. Then using the classical Morrey-Yosida approximation and suitable smoothing argument, a convex, smooth, bounded from below, and monotone increasing sequence ˜fn ((see [67] for details)) was used to regularize ˜f. Together with the Galerkin projection Pm:L2Hm:=span{φ1,φm}, where {φ1,φm,} is an orthonormal basis of L2(Ω) consisting eigenvectors of the Laplacian operator, a three-level approximation scheme was constructed:

    Qmt=Pm{δ˜FδQ+λI2+μμt}εPm{˜fnQ} (2.41)

    After obtaining the existence of global solutions for the approximation system (2.41), passing to limit with the order N, m, ε0+ led to the desired result.

    Theorem 2.10. There exists an explicitly computable constant η2 that depends on Li,i=1,,4 and Ω, such that if Q0H1(Ω)L(Ω), ˜QH32(Ω), and the smallness conditions in Lemma 2.2 are replaced by η2, then the system (2.38)-(2.39) admits a unique global weak solution under the coefficient assumption (2.40). Moreover, the initial smallness is preserved for all time.

    Under the same 2D settings for initial data with small L norm, the existence and uniqueness of global classical solutions of the Cauchy problem was established in [48], whose result was later improved in [55]. A stable numeric scheme and its convergence analysis were provided in [42], which provided an alternative way to prove the existence of global weak solutions to (2.38)-(2.39). The corresponding biaxial gradient flow in two dimensions was recently studied in [59], where global existence of weak and classical solutions were established under the assumption of small initial data. Concerning the co-rotational Beris-Edwards system whose elastic energy takes the form (2.33), it was proved in [62] that under periodic settings the hydrodynamic system admits a unique global weak solution if the initial L norm of the Q-tensor is properly small.

    Unsolved research problem 6: For the gradient flow (2.37) into the 3-dimensional Q-tensor space Q, prove the global wellposedness of the initial-boundary value problem under the assumption of small Q0L.

    The essential difference between the gradient flow (2.37) into the 2 and 3-dimensional Q-tensor spaces is that in the former case the system (2.38) can be written as

    Qijt=ζΔQijaQijctr(Q2)Qij+2L4k(QlklQij)L4iQkljQkl+L42|Q|2δij, (2.42)

    where

    ζ:=2L1+L2+L3>0

    due to coercivity condition (2.40) in 2D. As a consequence, a weak maximum principle argument can be used to prove Lemma 2.1 and henceforth the smallness of Q(t)L could be a priori kept during time evolution. In the latter case, however, the L1,L2,L3 terms in (2.38) cannot be combined into a single Laplace term ΔQ, so that one cannot establish a counterpart of Lemma 2.1 to ensure the smallness of Q(t)L for all time. In [48], a global wellposedness result was given with rather strong assumptions imposed on largeness of L1 and smallness of initial Q0 in terms of C2,α norm.

    On the other hand, it was shown in [28] that certain solutions exhibit a finite time blowup once the initial smallness assumption in Theorem 2.10 is violated.

    Theorem 2.11. There exists a smooth domain Ω, smooth initial data Q0, and a smooth function ˜Q:ΩR, under the coefficient assumption (2.40), the problem (2.38)-(2.39) does not admit a global smooth solution.

    The key ingredient in the proof of Theorem 2.11 was to show Q(t)L2(Ω) in finite time for any smooth solution. To this end, the following hedgehog type ansatz was used

    Qij(t,x)=θ(t,|x|)(xixj|x|2δij2),i,j=1,2.

    for

    Ω=BR1(0)BR0(0)R2.

    The initial data Q0 was assumed to be smooth and in the form of

    Q(0)=θ0(|x|)(x|x|x|x|I2),

    while the boundary data satisfies

    θ(t,R0)=θ(t,R1)0,  t>0.

    Then to show the finite time blowup of Q, it was equivalent to find the finite time blowup of θ that satisfies

    tθ=L4((θ)22+θθr+θθ+6θ2r2)+ζθ+ζθr4ζθr2aθcθ32. (2.43)

    To proceed, a non-linear differential inequality of the following quantity that blows up in finite time was established:

    12ddtR1R0θ2rdr2L4R0R41R40[π29(R1R0)21R20](R1R0θ2rdr)32|a|R1R0θ2rdr+4F(0),

    where

    F(t):=R1R0{L4θ[(θ)222θ2r2]ζ[(θ)22+2θ2r2](a2θ2+c8θ4)}rdrF(0),

    played a key role in controlling the high order nonlinearity even though the sign of the energy was not a priori known. Finally, the finite time blowup was ensured by assuming

    R20π29(R1R0)2>1.

    Unsolved research problem 7: For general domains which are not radially symmetric, determine whether the finite time blowup of solutions occurs for large enough initial data.

    This problem is much more difficult than the proof of Theorem 2.11 in that the problem (2.38)-(2.39) can no longer be reduced to a scalar equation. Further, boundary data has also to be dealt with in a subtle manner.

    To further overcome this issue of unboundedness from below caused by the cubic L4 term, motivated by the work in [56], a Maier-Saupe [64] type singular bulk potential was used in [44]. This potential can be traced back to the mean field theory.

    Nematic liquid crystals are a class of condensed matter systems whose rod-like molecules yield rich nonlinear phenomena, including isotropic-nematic phase transitions. Thermotropic nematic liquid crystals possess optical properties that change dramatically with variation of surrounding temperature. Above a certain temperature threshold their molecules are randomly oriented (that corresponds to the isotropic phase), whereas below this threshold they prefer to be aligned locally in a preferred direction (that relates to the nematic phase). The mean field theory, was originally proposed to describe such nematic-isotropic phase transition. In this theory, the liquid crystal molecular alignment was characterized by ρ(x,n), the density distribution function of the orientation of all molecules at a material point xΩR3. The de Gennes Q-tensor, which is defined as the deviation of the second moment of ρ from its isotropic value, is expressed as

    Q=S2[ρ(n)ρ(n)13I3]dσ(n). (2.44)

    It is indeed the normalized second-order moment of the probability measure ρ on S2, and it serves as an order parameter in the sense that it vanishes in the isotropic phase. Further, any de Gennes Q-tensor is symmetric, traceless, with all eigenvalues bounded between 1/3 and 2/3. Note that the cases of equality correspond to perfect crystalline nematic alignment, hence they are excluded [67]. Therefore, the eigenvalues of any de Gennes Q-tensor satisfy

    13<λi(Q)<23,1i3. (2.45)

    By convention, any QQ that satisfies the eigenvalue inequality (2.45) is called a physical Q-tensor, otherwise it is called unphysical. And we denote the space of physical Q-tensors by

    Qphy:={QQ|13<λi(Q)<113,1i3}. (2.46)

    Following the idea in [56], the singular potential fMS is defined by

    fMS(Q)def={infρAQS2ρ(n)lnρ(n)dn,13<λi(Q)<23,1i3+,otherwise, (2.47)

    where the admissible set AQ is

    AQ={ρP(S2),ρ(n)=ρ(n),S2[ρ(n)ρ(n)13I3]dn=Q} (2.48)

    Roughly speaking, we minimize the Boltzmann entropy over all probability distributions ρ with fixed normalized second moment QQphy, whose eigenvalues all stay within the interval (1/3,2/3). Correspondingly,

    ψB(Q)=fMS(Q)κ2|Q|2,κ>0, (2.49)

    is used to replace the regular polynomial bulk potential f, where the last term κ2|Q|2 is added to ensure the existence of local energy minimizers:

    ˜E:=[L1|Q|2+L2jQikkQij+L3jQijkQik+L4QlkkQijlQij+ψB(Q)]dx (2.50)

    Therefore, a natural enforcement of a physical constraint is imposed on the eigenvalues of the mathematical Q-tensor, and henceforth the free energy (2.33) could be kept under control under mild assumptions on L1,L4 [28,49,57].

    This Maier-Saupe type singular potential fMS(Q) is (see [10,52,66] for detailed proofs)

    ● isotropic, i.e.,

    fMS(Q)=fMS(BtQB),BSO(3).

    ● It is strictly convex.

    ● It is smooth in its effective domain D(fMS) where fMS assumes finite value, that is, when

    13<λi(Q)<23,1i3.

    Moreover, for each given QQphy, there exists a unique ρAQ satisfying

    fMS(Q)=S2ρQ(n)lnρQ(n)dσ(n). (2.51)

    And ρQ is given implicitly by

    ρQ(n)=exp(nn:μ)Z(μ),μQ, (2.52)

    where

    Z(μ)=S2exp(nn:μ)dσ(n) (2.53)

    satisfies

    1ZZμ=Q+13I3. (2.54)

    Due to the singular feature of fMS and the implicit formula to express fMS, the relevant analysis is rather challenging in both static and dynamic settings. For the free energy ˜E with L2=L3=L4=0, the associated Beris-Edwards system was studied analytically in [17,51,52,67]. More precisely, the existence of global weak solutions to a non-isothermal co-rotational Beris-Edwards syste was established in [51,52]. The existence, regularity and strict physicality of global weak solutions of the corresponding isothermal co-rotational Beris-Edwards system under periodic settings was investigated in [67]:

    {tu+(u)u+P=νΔu+(τ+σ),u=0,tQ+uQS(u,Q)=Γ[LΔQfMSQ+1dtr(fMSQ)I+κQ]. (2.55)

    Compared to the co-rotational Beris-Edwards system (2.1) in which ξ=0, the major difference in (2.55) is that the variational derivative of ˜E(Q) contains a singular term fMSQ other than regular polynomial terms. Besides the well-posedness result, one interesting topic to study is whether the hydrodynamic system (2.55) respects the physicality property as time evolves. To this end we shall define

    Definition 2.2. A map Q from a spatial domain Ω into Q is said to be strictly physical, if there exists δ>0 sufficiently small such that

    13+δλi(Q(x))23δ,1i3,a.e.xΩ. (2.56)

    Further, it is ready to check QQ is strictly physical if and only if fMS(Q)L, while fMS(Q)L1 ensures Q is physical. The main result in [67] states that

    Theorem 2.12. Suppose (u0,Q0)L2(Td)×H1(Td), u0=0, d=2,3, and

    fMS(Q0())L1(Td)(physical initial data),

    Then there exists a global weak solution (u,Q) of the system (2.55). Furthermore,

    fMS(Q(t,))L(Td),t>0(strictly physical at all positive time).

    The problem discussed in Theorem 2.12 is an important characteristic of evolution equations of parabolic type, bacause it is always natural to ask whether a solution with L1 initial data is instantaneously in L for all t>0. The essential idea in the proof of Theorem 2.12 was that although the evolution equation for the tensor field Q cannot be used to close an equation for the scalar function fMS(Q), a parabolic inequality satisfied by fMS(Q) can be derived so that a maximum principle argument led to strict physicality of weak solutions.

    Meanwhile, it was also proved in [67] that under 2D settings (for both the spatial dimension and the Q-tensor space) the existence of global strong solutions of the system (2.55) provided the initial data Q0 is strictly physical. Afterwards, the global existence and partial regularity results of a suitable weak solution to this system were proved in [17].

    For the free energy ˜E with L4=0, the relevant analysis is more challenging in both static and dynamic problems in that the appearance of anisotropic L2,L3 terms makes the maximum principle argument no longer be valid. Concerning static problems, the regularity results regarding energy minimizers under simpler settings of ˜F were discussed in [47,50,53], respectively. Concerning dynamic problems, the authors in [60] studied regularity properties of an L2 gradient flow generated by ˜E with L4=0 under periodic settings for d=2,3:

    {tQ(t,)=˜E(Q(t,)),t>0,Q(0,x)=Q0(x),xTd (2.57)

    More specifically, it was shown that for rather general initial data the associated L2 gradient flow admits a unique strong solution after any positive time, and this solution detaches from the physical boundary after a sufficiently large time T0.

    Theorem 2.13. For d=2,3 and any initial data

    Q0¯{QL2(Td;Qphy)|˜E(Q)<}L2(Td), (2.58)

    there exists a unique global solution Q(t,x):R+×TdQphy of (2.57) such that

    tQij=2L1ΔQij+2(L2+L3)jkQik23(L2+L3)klQlkδijfMSQij+13tr(fMSQ)δij+2κQij,a.e.(t,x)(0,)×Td. (2.59)

    Besides, Q(t,x) is physical for all t>0, a. e. xTd. Further, for any fixed t0>0, we have QL(t0,+;H2(Td)). Moreover, under the stronger assumption

    L13|L2+L3|κC2Td>0,

    where CTd=(2π)d is the Poincaré constant in Td, there exists T0>0 such that the unique solution is strictly physical for all tT0.

    The proof of pointwise existence of solutions in Theorem 2.13 relies on a powerful framework provided by Ambrosio-Gigli-Savare to obtain the solution of a gradient flow generated by a free energy (that is λ-convex) under very general assumptions of the initial data. However, there are essential difficulties caused by anisotropic L2+L3 terms to establish a uniform-in-time H2 bound of the solution Q. Without the L2+L3 terms in (2.59), the convexity of fMS together with classical L1L estimate of heat equation would ensure the strict physicality at any positive time and henceforth conventional energy method can be used. But the appearance of L2+L3 terms makes (2.59) a non-diagonal parabolic system with a singular potential which trends to infinity logarithmically when Q approaches its physical boundary, so that Q might not stay inside any compact subset of Qphy. To achieve the uniform-in-time H2 bound of Q, several results from the gradient flow theory by Ambrosio-Gigli-Savare as well as Gamma-convergence of gradient flows by Sandier and Serfaty were combined, and the gradient flow structure of equation (2.59) was carefully exploited.

    Furthermore, the estimate of the Hausdorff measure of the singular set where the solution touches the physical boundary in the intermediate stage (0,T0) was also provided in [60].

    Theorem 2.14. Let Q(t,x) be the unique strong solution of (2.57) established in Theorem 2.13. Then for a.e. t(0,T0), the contact set

    Σt:={xTd|Q(t,x)Qphy} (2.60)

    where Qphy denotes the boundary of Qphy, has the following estimate:

    dimH(Σt)2 for d=3.

    dimH(Σt)=0 for d=2.

    A crucial step in the proof of Theorem 2.14 was to establish the blowup rate of the gradient of fMS as Q approaches its physical boundary. The blowup rate estimates of fMS and fMS were initially studied in [43,45], while the latest were given in [63]:

    Theorem 2.15. For any QQphy, assume λ1(Q)λ2(Q)λ3(Q). Then the singular potential fMS is bounded above by

    fMS(Q)ln8312ln(λ1(Q)+13)12ln(λ2(Q)+13). (2.61)

    In the meantime, there exists a small computable constant δ0>0, whenever λ2(Q)+1/3<δ0, we have

    fMS(Q)ln168lnππ51612ln(λ1(Q)+13)12ln(λ2(Q)+13). (2.62)

    In addition, there exists a small computable constant ε0>0, whenever λ1(Q)+1/3<ε0, it holds

    C1λ1(Q)+13|QfMS(Q)|C2λ1(Q)+13, (2.63)

    where C1,C2>0 are explicitly computable constants and

    QfMS=fMSQ13tr(fMSQ)I3,

    It is noted that the blowup rate of fMS in Theorem 2.15 ranges from αln(λ1(Q)+1/3), 1/2α1, as Q approaches its physical boundary via various directions. The bounds (2.61) and (2.62) are of significance also because they imply

    fMS(Q)+12ln(λ1(Q)+13)+12ln(λ2(Q)+13)

    is a well defined, bounded function in the domain of λ1,λ2. Therefore, an accurate numerical approximation of fMS(Q) becomes possible by interpolating this function. On the other hand, the blowup rate estimate of QfMS plays a key role in both dynamic and static problems related to ˜E whenever anisotropic L2,L3 terms are present.

    Unsolved research problem 8: Prove the global-wellposeness of the L2 gradient flow generated by ˜E provided ˜E(Q0)<.

    Once L1,L4 terms are all present in the gradient flow, its relevant analysis becomes more challenging not only because of invalidity of maximum principle argument, but due to the loss of λ-convexity of the free energy ˜E as well.

    Finally, we want to point out that a relatively comprehensive list of references on the analytic study of static Q-tensor theory could be found in [68].

    The author would like to thank the anonymous referees for their useful suggestions to improve the quality of this paper. X. Xu's work is supported by the NSF grant DMS-2007157 and the Simons Foundation Grant No. 635288.

    The authors declare there is no conflicts of interest.



    [1] S. Chandrasekhar, Liquid Crystals, Cambridge U. Press, Cambridge, 1977.
    [2] P. G. de Gennes, J. Prost, The Physics of Liquid Crystals, Oxford Science Publications, Oxford, 1993.
    [3] J. Ericksen, Liquid crystals with variable degree of orientation, Arch Rational Mech. Anal., 113 (1990), 97–120. https://doi.org/10.1007/BF00380413 doi: 10.1007/BF00380413
    [4] E. Virga, Variational Theories for Liquid Crystals, Applied Mathematics and Mathematical Computation, 8, Chapman & Hall, London, 1994. https://doi.org/10.1007/978-1-4899-2867-2
    [5] F. C. Frank, On the theory of liquid crystals, Discuss. Faraday Soc., 25 (1958), 19–28.
    [6] N. J. Mottram, J. P. Newton, Introduction to Q-tensor theory, arXiv preprint, arXiv: 1409.3542, 2014.
    [7] R. Hardt, D. Kinderlehrer, F. H. Lin, Existence and partial regularity of static liquid crystal configurations, Comm. Math. Phys., 105 (1986), 547–570. https://doi.org/10.1007/BF01238933 doi: 10.1007/BF01238933
    [8] O. Alper, R. Hardt, F. H. Lin, Defects of liquid crystals with variable degree of orientation, Calc. Var. Partial Dif., 56 (2017), Paper No. 128. https://doi.org/10.1007/s00526-017-1218-5 doi: 10.1007/s00526-017-1218-5
    [9] F. H. Lin, On nematic liquid crystals with variable degree of orientation, Comm. Pure Appl. Math., 44 (1991), 453–468. https://doi.org/10.1002/cpa.3160440404 doi: 10.1002/cpa.3160440404
    [10] J. Ball, Mathematics of liquid crystals, Cambridge Centre for Analysis short course, (2012), 13–17.
    [11] A. Zarnescu, Topics in the Q-tensor theory of liquid crystals, Topics in mathematical modeling and analysis, 187–252, Jindrich Necas Cent. Math. Model. Lect. Notes, 7, Matfyzpress, Prague, 2012.
    [12] M. Paicu, A. Zarnescu, Global existence and regularity for the full coupled Navier-Stokes and Q-tensor system, SIAM J. Math. Anal., 43 (2011), 2009–2049. https://doi.org/10.1137/10079224X doi: 10.1137/10079224X
    [13] M. Paicu, A. Zarnescu, Energy dissipation and regularity for a coupled Navier-Stokes and Q-tensor system, Arch. Ration. Mech. Anal., 203 (2012), 45–67. https://doi.org/10.1007/s00205-011-0443-x doi: 10.1007/s00205-011-0443-x
    [14] A. N. Beris, B. J. Edwards, Thermodynamics of Flowing Systems with Internal Microstructure, Oxford Engineerin Science Series, No. 36, Oxford university Press, Oxford, New York, 1994.
    [15] F. De Anna, A global 2D well-posedness result on the order tensor liquid crystal theory, J. Differ. Equ., 262 (2017), 3932–3979. https://doi.org/10.1016/j.jde.2016.12.006 doi: 10.1016/j.jde.2016.12.006
    [16] M. M. Dai, E. Feireisl, E. Rocca, G. Schimperna, M. Schonbek, On asymptotic isotropy for a hydrodynamic model of liquid crystals, Asymptot. Anal., 97 (2016), 189–210. https://doi.org/10.3233/ASY-151348 doi: 10.3233/ASY-151348
    [17] H. R. Du, X. P. Hu, C. Y. Wang, Suitable weak solutions for the co-rotational Beris-Edwards system in dimension three, Arch. Ration. Mech. Anal., 238 (2020), 749–803. https://doi.org/10.1007/s00205-020-01554-y doi: 10.1007/s00205-020-01554-y
    [18] F. Guillén-González, M. A. Rodríguez-Bellido, Weak solutions for an initial-boundary Q-tensor problem related to liquid crystals, Nonlinear Anal., 112 (2015), 84–104. https://doi.org/10.1016/j.na.2014.09.011 doi: 10.1016/j.na.2014.09.011
    [19] H. Abels, G. Dolzmann, Y. N. Liu, Strong solutions for the Beris–Edwards model for nematic liquid crystals with homogeneous Dirichlet boundary conditions, Adv. Differ. Equ., 21 (2016), 109–152.
    [20] F. Guillén-González, M. A. Rodríguez-Bellido, Weak time regularity and uniqueness for a Q-tensor model, SIAM J. Math. Anal., 46 (2014), 3540–3567. https://doi.org/10.1137/13095015X doi: 10.1137/13095015X
    [21] F. De Anna, A. Zarnescu, Uniqueness of weak solutions of the full coupled Navier–Stokes and Q-tensor system in 2D, Commun. Math. Sci., 14 (2016), 2127–2178. https://doi.org/10.4310/CMS.2016.v14.n8.a3 doi: 10.4310/CMS.2016.v14.n8.a3
    [22] C. Cavaterra, E. Rocca, H. Wu, X. Xu, Global strong solutions of the full Navier–Stokes and Q-tensor system for nematic liquid crystal flows in two dimensions, SIAM J. Math. Anal., 48 (2016), 1368–1399. https://doi.org/10.1137/15M1048550 doi: 10.1137/15M1048550
    [23] H. Abels, G. Dolzmann, Y. N. Liu, Well-posedness of a fully coupled Navier–Stokes/Q-tensor system with inhomogeneous boundary data, SIAM J. Math. Anal., 46 (2014), 3050–3077. https://doi.org/10.1137/130945405 doi: 10.1137/130945405
    [24] J. Ericksen, Conservation laws for liquid crystals, Trans. Soc. Rheol., 5 (1961), 22–34. https://doi.org/10.1122/1.548883 doi: 10.1122/1.548883
    [25] F. Leslie, Some constitutive equations for liquid crystals, Arch. Ration. Mech. Anal., 28 (1968), 265–283. https://doi.org/10.1007/BF00251810 doi: 10.1007/BF00251810
    [26] W. Wang, P. W. Zhang, Z. F. Zhang, Rigorous derivation from Landau–de Gennes theory to Ericksen–Leslie theory, SIAM J. Math. Anal., 47 (2015), 127–158. https://doi.org/10.1137/13093529X doi: 10.1137/13093529X
    [27] F. H. Lin, C. Y. Wang, Recent developments of analysis for hydrodynamic flow of nematic liquid crystals, Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 372 (2014), 20130361. https://doi.org/10.1098/rsta.2013.0361 doi: 10.1098/rsta.2013.0361
    [28] G. Iyer, X. Xu, A. Zarnescu, Dynamic cubic instability in a 2D Q-tensor model for liquid crystals, Math. Mod. Methods Appl. Sci., 25 (2015), 1477–1517. https://doi.org/10.1142/S0218202515500396 doi: 10.1142/S0218202515500396
    [29] H. Wu, X. Xu, A. Zarnescu, Dynamics and flow effects in the Beris-Edwards system modelling nematic liquid crystals, Arch. Rational Mech. Anal., 231 (2019), 1217–1267. https://doi.org/10.1007/s00205-018-1297-2 doi: 10.1007/s00205-018-1297-2
    [30] A. Contreras, X. Xu, W. J. Zhang, An elementary proof of eigenvalue preservation for the co-rotational Beris-Edwards system, J. Nonlinear Sci., 29 (2019), 789–801. https://doi.org/10.1007/s00332-018-9503-9 doi: 10.1007/s00332-018-9503-9
    [31] A. C. Murza, A. E. Teruel, A. Zarnescu, Shear flow dynamics in the Beris-Edwards model of nematic liquid crystals, Proc. R. Soc. A, 474 (2018), 20170673. https://doi.org/10.1098/rspa.2017.0673 doi: 10.1098/rspa.2017.0673
    [32] T. Qian, P. Sheng, Generalized hydrodynamic equations for nematic liquid crystals, Phys. Rev. E, 58 (1998), 7475. https://doi.org/10.1103/PhysRevE.58.7475 doi: 10.1103/PhysRevE.58.7475
    [33] F. Leslie, Theory of flow phenomena in liquid crystals, Adv. Liquid Crystals, 4 (1979), 1–81. https://doi.org/10.1016/B978-0-12-025004-2.50008-9 doi: 10.1016/B978-0-12-025004-2.50008-9
    [34] S. R. Li, W. Wang, Rigorous justification of the uniaxial limit from the Qian-Sheng inertial Q-tensor theory to the Ericksen-Leslie theory, SIAM J. Math. Anal., 52 (2020), 4421–4468. https://doi.org/10.1137/19M129200X doi: 10.1137/19M129200X
    [35] F. De Anna, A. Zarnescu, Global well-posedness and twist-wave solutions for the inertial Qian-Sheng model of liquid crystals, J. Differ. Equ., 264 (2018), 1080–1118.
    [36] J. Ericksen, Twisted waves in liquid crystals, Q. J. Mech. Appl. Math., 21 (1968), 463–465. https://doi.org/10.1093/qjmam/21.4.463 doi: 10.1093/qjmam/21.4.463
    [37] N. Jiang, Y. L. Luo, Y. Ma, S. J. Tang, Entropy inequality and energy dissipation of inertial Qian-Sheng model for nematic liquid crystals, J. Hyperbolic Differ. Equ., 18 (2021), 221–256. https://doi.org/10.1142/S0219891621500065 doi: 10.1142/S0219891621500065
    [38] E. Feireisl, E. Rocca, G. Schimperna, A. Zarnescu, On a hyperbolic system arising in liquid crystals modeling, J. Hyperbolic Differ. Equ., 15 (2018), 15–35. https://doi.org/10.1142/S0219891618500029 doi: 10.1142/S0219891618500029
    [39] G. Chen, T. Huang, W. S. Liu, Poiseuille flow of nematic liquid crystals via the full Ericksen-Leslie model, Arch. Ration. Mech. Anal., 236 (2020), 839–891. https://doi.org/10.1007/s00205-019-01484-4 doi: 10.1007/s00205-019-01484-4
    [40] M. W. Fei, W. Wang, P. W. Zhang, Z. F. Zhang, On the isotropic-nematic phase transition for the liquid crystal, Peking Math. J., 1 (2018), 141–219. https://doi.org/10.1007/s42543-018-0005-3 doi: 10.1007/s42543-018-0005-3
    [41] T. Laux, Y. N. Liu, Nematic-Isotropic Phase Transition in Liquid Crystals: A Variational Derivation of Effective Geometric Motions, Arch. Ration. Mech. Anal., 241 (2021), 1785–1814. https://doi.org/10.1007/s00205-021-01681-0 doi: 10.1007/s00205-021-01681-0
    [42] Y. Y. Cai, J. Shen, X. Xu, A stable scheme and its convergence analysis for a 2D dynamic Q-tensor model of nematic liquid crystals, Math. Mod. Methods Appl. Sci., 27 (2017), 1459–1488. https://doi.org/10.1142/S0218202517500245 doi: 10.1142/S0218202517500245
    [43] J. Ball, Analysis of liquid crystals and their defects, Lecture notes of the Scuola Estiva GNFM, Ravello 17–22 Sep. 2018.
    [44] J. Ball, A. Majumdar, Nematic liquid crystals: from Maier-Saupe to a continuum theory, Mol. Cryst. Liq. Cryst., 525 (2010), 1–11. https://doi.org/10.1080/15421401003795555 doi: 10.1080/15421401003795555
    [45] J. Ball, A. Majumdar, Passage from the mean-field Maier-Saupe to the continuum Landau-de Gennes theory for nematic liquid crystals, work in progress.
    [46] J. Ball, A. Zarnescu, Orientability and energy minimization in liquid crystal models, Arch. Ration. Mech. Anal., 202 (2011), 493–535. https://doi.org/10.1007/s00205-011-0421-3 doi: 10.1007/s00205-011-0421-3
    [47] P. Bauman, D. Phillips, Regularity and the behavior of eigenvalues for minimizers of a constrained Q-tensor energy for liquid crystals, Calc. Var. Partial Dif., 55 (2016), 55–81. https://doi.org/10.1007/s00526-016-1009-4 doi: 10.1007/s00526-016-1009-4
    [48] X. F. Chen, X. Xu, Existence and uniqueness of global classical solutions of a gradient flow of the Landau–de Gennes energy, Proc. Amer. Math. Soc., 144 (2016), 1251–1263. https://doi.org/10.1090/proc/12803 doi: 10.1090/proc/12803
    [49] T. A. Davis, E. C. Gartland, Finite element analysis of the Landau-de Gennes minimization problem for liquid crystals, SIAM J. Numer. Anal., 35 (1998), 336–362. https://doi.org/10.1137/S0036142996297448 doi: 10.1137/S0036142996297448
    [50] L. C. Evans, O. Kneuss, H. Tran, Partial regularity for minimizers of singular energy functionals, with application to liquid crystal models, Trans. Amer. Math. Soc., 368 (2016), 3389–3413.
    [51] E. Feireisl, E. Rocca, G. Schimperna, A. Zarnescu, Evolution of non-isothermal Landau-de Gennes nematic liquid crystals flows with singular potential, Commun. Math. Sci., 12 (2014), 317–343. https://doi.org/10.4310/CMS.2014.v12.n2.a6 doi: 10.4310/CMS.2014.v12.n2.a6
    [52] E. Feireisl, E. Rocca, G. Schimperna, A. Zarnescu, Nonisothermal nematic liquid crystal flows with the Ball-Majumdar free energy, Ann. di Mat. Pura ed Appl., 194 (2015), 1269–1299. https://doi.org/10.1007/s10231-014-0419-1 doi: 10.1007/s10231-014-0419-1
    [53] Z. Y. Geng, J. J. Tong, Regularity of minimizers of a tensor-valued variational obstacle problem in three dimensions, Calc. Var., 59 (2020), Paper No. 57. https://doi.org/10.1007/s00526-020-1717-7 doi: 10.1007/s00526-020-1717-7
    [54] J. R. Huang, S. J. Ding, Global well-posedness for the dynamical Q-tensor model of liquid crystals, Sci. China Math., 58 (2015), 1349–1366. https://doi.org/10.1007/s11425-015-4990-8 doi: 10.1007/s11425-015-4990-8
    [55] T. Huang, N. Zhao, On the regularity of weak small solution of a gradient flow of the Landau–de Gennes energy, Proc. Amer. Math. Soc., 147 (2019), 1687–1698. https://doi.org/10.1090/proc/14337 doi: 10.1090/proc/14337
    [56] J. Katriel, G. Kventsel, G. Luckhurst, T. Sluckin, Free energies in the Landau and molecular field approaches, Liq. Cryst., 1 (1986), 337–355. https://doi.org/10.1080/02678298608086667 doi: 10.1080/02678298608086667
    [57] G. Kitavtsev, J. M. Robbins, V. Slastikov, A. Zarnescu, Liquid crystal defects in the Landau-de Gennes theory in two dimensions-beyond the one-constant approximation, Math. Mod. Methods Appl. Sci., 26 (2016), 2769–2808. https://doi.org/10.1142/S0218202516500664 doi: 10.1142/S0218202516500664
    [58] F. H. Lin, C. Poon, On Ericksen's model for liquid crystals, J. Geom. Anal., 4 (1994), 379–392. https://doi.org/10.1007/BF02921587 doi: 10.1007/BF02921587
    [59] J. Y. Lin, C. Zhou, Existence of solutions to the biaxial nematic liquid crystals with two order parameter tensors, Math. Methods Appl. Sci., 43 (2020), 6430–6453. https://doi.org/10.1002/mma.6387 doi: 10.1002/mma.6387
    [60] Y. N. Liu, X. Y. Lu, X. Xu, Regularity of a gradient flow generated by the anisotropic Landau-de Gennes energy with a singular potential, SIAM J. Math. Anal., 53 (2021), 3338–3365. https://doi.org/10.1137/20M1386499 doi: 10.1137/20M1386499
    [61] Y. N. Liu, W. Wang, On the initial boundary value problem of a Navier-Stokes/Q-tensor model for liquid crystals, Discrete Contin. Dyn. Syst. Ser. B, 23 (2018), 3879–3899. https://doi.org/10.3934/dcdsb.2018115 doi: 10.3934/dcdsb.2018115
    [62] Y. N. Liu, H. Wu, X. Xu, Global well-posedness of the two dimensional Beris-Edwards system with general Laudau-de Gennes free energy, J. Differ. Equ., 267 (2019), 6958–7001. https://doi.org/10.1016/j.jde.2019.07.010 doi: 10.1016/j.jde.2019.07.010
    [63] X. Y. Lu, X. Xu, W. J. Zhang, Blowup rate estimates of a singular potential and its gradient in the Landau-de Gennes theory, Journal of Nonlinear Sci., 32 (2022), 1–30. https://doi.org/10.1007/s00332-021-09761-x doi: 10.1007/s00332-021-09761-x
    [64] W. Maier, A. Saupe, A simple molecular statistical theory of the nematic crystalline-liquid phase, I Z Naturf. a, 14 (1959), 882–889.
    [65] A. Majumdar, A. Zarnescu, Landau-De Gennes theory of nematic liquid crystals: the Oseen-Frank limit and beyond, Arch. Ration. Mech. Anal., 196 (2010), 227–280. https://doi.org/10.1007/s00205-009-0249-2 doi: 10.1007/s00205-009-0249-2
    [66] C. D. Schimming, J. Vinals, S. W. Walker, Numerical method for the equilibrium configurations of a Maier-Saupe bulk potential in a Q-tensor model of an anisotropic nematic liquid crystal, J. Comput. Phys., 441 (2021), Paper No. 110441. https://doi.org/10.1016/j.jcp.2021.110441 doi: 10.1016/j.jcp.2021.110441
    [67] M. Wilkinson, Strict physicality of global weak solutions of a Navier-Stokes Q-tensor system with singular potential, Arch. Ration. Mech. Anal., 218 (2015), 487–526. https://doi.org/10.1007/s00205-015-0864-z doi: 10.1007/s00205-015-0864-z
    [68] A. Zarnescu, Mathematical problems of nematic liquid crystals: between dynamical and stationary problems, Trans. R. Soc. Lond. Ser. A., 379 (2021), Paper No. 20200432. https://doi.org/10.1098/rsta.2020.0432 doi: 10.1098/rsta.2020.0432
  • This article has been cited by:

    1. Justin Swain, Giordano Tierra, Linear numerical schemes for a Q-tensor system for nematic liquid crystals, 2024, 430, 00457825, 117190, 10.1016/j.cma.2024.117190
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2340) PDF downloads(171) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog