This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Variational and numerical analysis of a 𝐐\mathbf{Q}-tensor model for smectic-A liquid crystals

Jingmin Xia [email protected] College of Meteorology and Oceanography, National University of Defense Technology, Changsha, China;  and  Patrick E. Farrell [email protected] Mathematical Institute, University of Oxford, Oxford, UK;
(Date: The dates will be set by the publisher)
Abstract.

We analyse an energy minimisation problem recently proposed for modelling smectic-A liquid crystals. The optimality conditions give a coupled nonlinear system of partial differential equations, with a second-order equation for the tensor-valued nematic order parameter 𝐐\mathbf{Q} and a fourth-order equation for the scalar-valued smectic density variation uu. Our two main results are a proof of the existence of solutions to the minimisation problem, and the derivation of a priori error estimates for its discretisation of the decoupled case (i.e., q=0q=0) using the 𝒞0\mathcal{C}^{0} interior penalty method. More specifically, optimal rates in the H1H^{1} and L2L^{2} norms are obtained for 𝐐\mathbf{Q}, while optimal rates in a mesh-dependent norm and L2L^{2} norm are obtained for uu. Numerical experiments confirm the rates of convergence.

Key words and phrases:
𝒞0\mathcal{C}^{0} interior penalty method, a priori error estimates, finite element methods, smectic liquid crystals
1991 Mathematics Subject Classification:
76A15, 49J10, 35B45, 65N12, 65N30
The work of JX is supported by the National Natural Science Foundation of China (No. 12201636) and the Research Fund of National University of Defense Technology [grant number ZK22-37]. The work of PEF is supported by EPSRC [grant numbers EP/R029423/1 and EP/W026163/1].

Introduction

Smectic liquid crystals (LC) are layered mesophases that have a periodic modulation of the mass density along one spatial direction. Informally, they can be thought of as one-dimensional solids along the direction of periodicity and two-dimensional fluids along the other two remaining directions. According to different in-layer structures, several types of smectic LC are recognised, e.g., smectic-A, smectic-B and smectic-C (see [26, Figure 9] for illustrations of different smectic phases). In particular, in the smectic-A phase, the molecules tend to align parallel to the normals of the smectic layers. For a broader review of liquid crystals, see [2, 16, 28].

Several models have been proposed to describe smectic-A liquid crystals. The classical de Gennes model [15] employs a complex-valued order parameter to describe the magnitude and phase of the density variation. This complex-valued parameterisation leads to some key modelling difficulties, which motivated the development by Pevnyi, Selinger & Sluckin (PSS) of a smectic-A model with a real-valued smectic order parameter and a vector-valued nematic order parameter [24]. The use of a vector-valued nematic order parameter limits the kinds of defects the model can permit [2], and hence Ball & Bedford (BB) proposed a version of the PSS model employing a tensor-valued nematic order parameter [3] instead. The model considered in this work is similar to the BB model, but with additional modifications to render it amenable to numerical simulation (see [31] for details). The model was used to numerically simulate several key smectic defect structures, such as oily streaks and focal conic domains, for the first time.

While the numerical modelling of nematic liquid crystals is now mature, relatively little work has considered smectics. Garcìa-Cervera and Joo [19] perform numerical simulations of the classical de Gennes model [15] in the presence of a magnetic field, using a combined Fourier-finite difference approach. Wittmann et al. use density functional theory to investigate the topology of smectic liquid crystals in complex confinement [30]. Monderkamp et al. examine the topology of defects in two-dimensional confined smectics with the help of extensive Monte Carlo simulations [22]. Our goal in this work is to analyse a model for smectic-A liquid crystals, and its finite element discretization, that was recently proposed by Xia, MacLachlan, Atherton and Farrell in [31].

We consider an open, bounded and convex spatial domain Ωd\Omega\subset\mathbb{R}^{d}, d{2,3}d\in\{2,3\} with Lipschitz boundary Ω\partial\Omega. The smectic-A free energy we analyse in this work is given by

𝒥(u,𝐐)=Ω(fs(u)+B|𝒟2u+q2(𝐐+𝐈dd)u|2+fn(𝐐,𝐐)),\mathcal{J}(u,\mathbf{Q})=\int_{\Omega}\left(f_{s}(u)+B\left|\mathcal{D}^{2}u+q^{2}\left(\mathbf{Q}+\frac{\mathbf{I}_{d}}{d}\right)u\right|^{2}+f_{n}(\mathbf{Q},\nabla\mathbf{Q})\right), (1)

where 𝐈d\mathbf{I}_{d} denotes the d×dd\times d identity matrix, 𝒟2\mathcal{D}^{2} is the Hessian operator, fs(u)f_{s}(u) is the smectic bulk energy density

fs(u)a12u2+a23u3+a34u4f_{s}(u)\coloneqq\frac{a_{1}}{2}u^{2}+\frac{a_{2}}{3}u^{3}+\frac{a_{3}}{4}u^{4}

and fn(𝐐,𝐐)f_{n}(\mathbf{Q},\nabla\mathbf{Q}) is the nematic bulk energy density

fn(𝐐,𝐐)\displaystyle f_{n}(\mathbf{Q},\nabla\mathbf{Q}) =fne(𝐐)+fnb(𝐐)\displaystyle=f_{n}^{e}(\nabla\mathbf{Q})+f_{n}^{b}(\mathbf{Q}) (2)
K2|𝐐|2+{(l(tr(𝐐2))+l(tr(𝐐2))2),if d=2,(l2(tr(𝐐2))l3(tr(𝐐3))+l2(tr(𝐐2))2),if d=3.\displaystyle\coloneqq\frac{K}{2}|\nabla\mathbf{Q}|^{2}+\begin{cases}\left(-l\left(\text{tr}(\mathbf{Q}^{2})\right)+l\left(\text{tr}(\mathbf{Q}^{2})\right)^{2}\right),&\text{if }d=2,\\ \left(-\frac{l}{2}\left(\text{tr}(\mathbf{Q}^{2})\right)-\frac{l}{3}\left(\text{tr}(\mathbf{Q}^{3})\right)+\frac{l}{2}\left(\text{tr}(\mathbf{Q}^{2})\right)^{2}\right),&\text{if }d=3.\end{cases}

Here, B>0B>0 is the nematic-smectic coupling parameter, a1,a2,a3a_{1},a_{2},a_{3} represent smectic bulk constants with a3>0a_{3}>0, K>0K>0 and l>0l>0 are nematic elastic and bulk constants, respectively, q0q\geq 0 is the wave number of the smectic layers, and the trace of a matrix is given by tr()\mathrm{tr}(\cdot).

The two main contributions of this paper are to prove existence of minimisers to the problem of minimising 𝒥\mathcal{J} over a suitably-defined admissible set, and to derive a priori error estimates for its discretisation using the 𝒞0\mathcal{C}^{0} interior penalty method [7, 9]. We show the existence result of minimising the free energy eq. 1 in section 1. We then derive a priori error estimates for both 𝐐\mathbf{Q} and uu in the simplified case q=0q=0 in section 2. We confirm that the theoretical predictions match numerical experiments in section 3, for both q=0q=0 and q>0q>0.

In order to avoid the proliferation of constants throughout this work, we use the notation aba\lesssim b (respectively, bab\gtrsim a) to represent the relation aCba\leq Cb (respectively, bCab\geq Ca) for some generic constant CC (possibly not the same constant on each appearance) independent of the mesh.

1. Existence of minimisers

To formulate the minimisation problem for eq. 1, we must first define the admissible space 𝒜\mathcal{A} in which minimisers will be sought. We define 𝒜\mathcal{A} as

𝒜={(u,𝐐)H2(Ω,)×H1(Ω,S0):𝐐=𝐐b,u=ubonΩ},\mathcal{A}=\bigg{\{}(u,\mathbf{Q})\in H^{2}(\Omega,\mathbb{R})\times H^{1}(\Omega,S_{0}):\mathbf{Q}=\mathbf{Q}_{b},u=u_{b}\ \text{on}\ \partial\Omega\bigg{\}}, (3)

with the specified Dirichlet boundary data 𝐐bH1/2(Ω,S0)\mathbf{Q}_{b}\in H^{1/2}(\partial\Omega,S_{0}) and ubH3/2(Ω,)u_{b}\in H^{3/2}(\partial\Omega,\mathbb{R}). Here, S0S_{0} denotes the space of symmetric and traceless d×dd\times d real-valued matrices. For simplicity of the analysis, we only consider Dirichlet boundary conditions for 𝐐\mathbf{Q} and uu here, but other types of boundary conditions (e.g., a mixture of the Dirichlet and natural boundary conditions for uu as illustrated in the implementations in [31]) can be taken. With this admissible space, we consider the minimisation problem

min(u,𝐐)𝒜𝒥(u,𝐐).\min_{(u,\mathbf{Q})\in\mathcal{A}}\mathcal{J}(u,\mathbf{Q}). (4)

Notice that fn(𝐐,𝐐)f_{n}(\mathbf{Q},\nabla\mathbf{Q}) is the classical Landau de Gennes (LdG) free energy density [16, 23] for nematic LC and it is proven by Davis & Gartland [14, Corollary 4.4] that there exists a minimiser of the functional Ωfn(𝐐,𝐐)\int_{\Omega}f_{n}(\mathbf{Q},\nabla\mathbf{Q}) for 𝐐H1(Ω,S0)\mathbf{Q}\in H^{1}(\Omega,S_{0}) in three dimensions. Moreover, Bedford [4, Theorem 5.18] gives an existence result for the BB model in three dimensions:

min(u,𝐐)𝒜BB𝒥BB(u,𝐐)=Ω{K2|𝐐|2+B|𝒟2u+q2(𝐐s+𝐈33)u|2+a12u2+a23u3+a34u4},\min_{(u,\mathbf{Q})\in\mathcal{A}^{BB}}\mathcal{J}^{BB}(u,\mathbf{Q})=\int_{\Omega}\left\{\frac{K}{2}|\nabla\mathbf{Q}|^{2}+B\left|\mathcal{D}^{2}u+q^{2}\left(\frac{\mathbf{Q}}{s}+\frac{\mathbf{I}_{3}}{3}\right)u\right|^{2}+\frac{a_{1}}{2}u^{2}+\frac{a_{2}}{3}u^{3}+\frac{a_{3}}{4}u^{4}\right\},

with an admissible space 𝒜BB{𝐐SBV(Ω,S0),uH2(Ω,):𝐐=s(𝐧𝐧𝐈dd),s[0,1],|𝐧|=1}\mathcal{A}^{BB}\coloneqq\{\mathbf{Q}\in SBV(\Omega,S_{0}),u\in H^{2}(\Omega,\mathbb{R}):\mathbf{Q}=s\left(\mathbf{n}\otimes\mathbf{n}-\frac{\mathbf{I}_{d}}{d}\right),s\in[0,1],|\mathbf{n}|=1\}, where SBVSBV denotes special functions of bounded variation. For simplicity, we have ignored the surface integral here in the energy functional of the BB model. One can observe its resemblance to eq. 1. Motivated by the above results, we prove the existence of minimisers to eq. 1 via the direct method of calculus of variations.

Theorem 1.1.

(Existence of minimisers) Let 𝒥\mathcal{J} be of the form eq. 1 with positive parameters a3a_{3}, BB, KK, ll and non-negative wave number qq. Then there exists a solution pair (u,𝐐)(u^{\star},\mathbf{Q}^{\star}) that solves eq. 4.

Proof.

Note that both the smectic density fsf_{s} and the nematic bulk density fnbf_{n}^{b} are bounded from below as a3,l>0a_{3},l>0. Thus, 𝒥\mathcal{J} is also bounded from below and we can choose a minimising sequence {(uj,𝐐j)}\{(u_{j},\mathbf{Q}_{j})\}, i.e.,

(uj,𝐐j)𝒜,uju~H2H01(Ω,),\displaystyle(u_{j},\mathbf{Q}_{j})\in\mathcal{A},\ u_{j}-\tilde{u}\in H^{2}\cap H^{1}_{0}(\Omega,\mathbb{R}), 𝐐j𝐐~H01(Ω,S0),\displaystyle\mathbf{Q}_{j}-\tilde{\mathbf{Q}}\in H^{1}_{0}(\Omega,S_{0}), (5)
𝒥(uj,𝐐j)jinf{𝒥(u,𝐐):(u,𝐐)𝒜,uu~H2H01(Ω,),\displaystyle\mathcal{J}(u_{j},\mathbf{Q}_{j})\overset{j\to\infty}{\longrightarrow}\inf\{\mathcal{J}(u,\mathbf{Q}):(u,\mathbf{Q})\in\mathcal{A},\ u-\tilde{u}\in H^{2}\cap H^{1}_{0}(\Omega,\mathbb{R}), 𝐐𝐐~H01(Ω,S0)}<.\displaystyle\mathbf{Q}-\tilde{\mathbf{Q}}\in H^{1}_{0}(\Omega,S_{0})\}<\infty.

Here we set 𝐐~H1(Ω,S0)\tilde{\mathbf{Q}}\in H^{1}(\Omega,S_{0}) (resp. u~H2(Ω,)\tilde{u}\in H^{2}(\Omega,\mathbb{R})) to be any function with trace 𝐐b\mathbf{Q}_{b} (resp. ubu_{b}). We tackle the three terms in 𝒥(u,𝐐)\mathcal{J}(u,\mathbf{Q}) separately in the following.

First, for the nematic energy term Ωfn(𝐐,𝐐)\int_{\Omega}f_{n}(\mathbf{Q},\nabla\mathbf{Q}), we can follow the proof of [14, Theorem 4.3] to obtain that fn(𝐐j,𝐐j)f_{n}(\mathbf{Q}_{j},\nabla\mathbf{Q}_{j}) is coercive in H1(Ω,S0)H^{1}(\Omega,S_{0}), i.e., fnf_{n} grows unbounded as 𝐐j1\|\mathbf{Q}_{j}\|_{1}\rightarrow\infty, and thus the minimising sequence {𝐐j}\{\mathbf{Q}_{j}\} must be bounded in 𝐇1(Ω,S0)\mathbf{H}^{1}(\Omega,S_{0}). Since H1(Ω)H^{1}(\Omega) is a reflexive Banach Space, we have a subsequence (also denoted as {𝐐j}\{\mathbf{Q}_{j}\}) that weakly converges to 𝐐H1(Ω,S0)\mathbf{Q}^{\star}\in H^{1}(\Omega,S_{0}) such that 𝐐𝐐~H01(Ω,S0)\mathbf{Q}^{\star}-\tilde{\mathbf{Q}}\in H^{1}_{0}(\Omega,S_{0}), and from the Rellich–Kondrachov theorem it follows that 𝐐j𝐐 in L2(Ω) and 𝐐j𝐐 in L2(Ω).\mathbf{Q}_{j}\rightarrow\mathbf{Q}^{\star}\text{ in }L^{2}(\Omega)\text{ and }\nabla\mathbf{Q}_{j}\rightharpoonup\nabla\mathbf{Q}^{\star}\text{ in }L^{2}(\Omega). The weak lower semi-continuity of the nematic energy density fnf_{n} in eq. 2 is guaranteed by [14, Lemma 4.2], therefore,

lim infjΩfn(𝐐j,𝐐j)Ωfn(𝐐,𝐐).\liminf_{j\to\infty}\int_{\Omega}f_{n}(\mathbf{Q}_{j},\nabla\mathbf{Q}_{j})\geq\int_{\Omega}f_{n}(\mathbf{Q}^{\star},\nabla\mathbf{Q}^{\star}). (6)

For the smectic bulk term Ωfs(u)\int_{\Omega}f_{s}(u), we can follow the proof in [4, Theorem 5.19]. By eq. 5, we have

supjΩ(|𝒟2uj|2+|uj|2)<,\sup_{j}\int_{\Omega}\left(\left|\mathcal{D}^{2}u_{j}\right|^{2}+|u_{j}|^{2}\right)<\infty,

which implies an upper bound for uj\nabla u_{j} using [4, Ineq. (5.42)]. Hence, {uj}\{u_{j}\} is bounded in H2(Ω)H^{2}(\Omega) and thus there is a subsequence (also denoted as {uj}\{u_{j}\}) such that uju in H2(Ω)u_{j}\rightharpoonup u^{\star}\text{ in }H^{2}(\Omega) and uu~H2H01(Ω)u^{\star}-\tilde{u}\in H^{2}\cap H^{1}_{0}(\Omega). Moreover, one can readily check that u<\|u^{\star}\|_{\infty}<\infty by the Sobolev embedding of H2(Ω)H^{2}(\Omega) into the Hölder spaces 𝒞𝔱,ϰ0(Ω)\mathcal{C}^{\mathfrak{t},\varkappa_{0}}(\Omega) (𝔱+ϰ0=1\mathfrak{t}+\varkappa_{0}=1 for d=2d=2 and 𝔱+ϰ0=1/2\mathfrak{t}+\varkappa_{0}=1/2 for d=3d=3) and the boundedness of domain Ω\Omega. Again, it follows from the Rellich–Kondrachov theorem that uju in L2(Ω) and 𝒟2uj𝒟2u in L2(Ω).u_{j}\rightarrow u^{\star}\text{ in }L^{2}(\Omega)\text{ and }\mathcal{D}^{2}u_{j}\rightharpoonup\mathcal{D}^{2}u^{\star}\text{ in }L^{2}(\Omega). Noting fs(u)f_{s}(u) is bounded from below for all uH2(Ω)u\in H^{2}(\Omega), we then obtain

lim infjΩfs(uj)Ωfs(u).\liminf_{j\to\infty}\int_{\Omega}f_{s}(u_{j})\geq\int_{\Omega}f_{s}(u^{\star}). (7)

Now, we consider the nematic-smectic coupling term ΩB|𝒟2u+q2(𝐐+𝐈dd)u|2\int_{\Omega}B\left|\mathcal{D}^{2}u+q^{2}\left(\mathbf{Q}+\frac{\mathbf{I}_{d}}{d}\right)u\right|^{2} in 𝒥(u,𝐐)\mathcal{J}(u,\mathbf{Q}). Note that when the wave number q=0q=0, this term reduces to ΩB|𝒟2u|2\int_{\Omega}B|\mathcal{D}^{2}u|^{2} and it is straightforward to obtain the weak lower semi-continuity property. Therefore, we discuss the case of q>0q>0 in detail as follows. By the 𝐇1\mathbf{H}^{1}-boundedness property of the minimising sequence {𝐐j}\{\mathbf{Q}_{j}\} and the fact that u<\|u^{\star}\|_{\infty}<\infty, we can deduce

Ω|uj𝐐ju𝐐|2\displaystyle\int_{\Omega}|u_{j}\mathbf{Q}_{j}-u^{\star}\mathbf{Q}^{\star}|^{2} =Ω|(uju)𝐐j+u(𝐐j𝐐)|2\displaystyle=\int_{\Omega}\left|(u_{j}-u^{\star})\mathbf{Q}_{j}+u^{\star}(\mathbf{Q}_{j}-\mathbf{Q}^{\star})\right|^{2}
2Ω(|uju|2|𝐐j|2+|u|2|𝐐j𝐐|2)\displaystyle\leq 2\int_{\Omega}\left(|u_{j}-u^{\star}|^{2}|\mathbf{Q}_{j}|^{2}+|u^{\star}|^{2}|\mathbf{Q}_{j}-\mathbf{Q}^{\star}|^{2}\right)
0as uju,𝐐j𝐐 in L2.\displaystyle\to 0\quad\text{as }u_{j}\to u^{\star},\mathbf{Q}_{j}\to\mathbf{Q}^{\star}\text{ in }L^{2}.

Hence, uj𝐐ju𝐐u_{j}\mathbf{Q}_{j}\to u^{\star}\mathbf{Q}^{\star} in L2(Ω)L^{2}(\Omega), and further,

uj(𝐐j+𝐈dd)\displaystyle u_{j}\left(\mathbf{Q}_{j}+\frac{\mathbf{I}_{d}}{d}\right) u(𝐐+𝐈dd)\displaystyle\to u^{\star}\left(\mathbf{Q}^{\star}+\frac{\mathbf{I}_{d}}{d}\right) in L2(Ω),\displaystyle\quad\text{in }L^{2}(\Omega),
uj(𝐐j+𝐈dd):𝒟2uj\displaystyle u_{j}\left(\mathbf{Q}_{j}+\frac{\mathbf{I}_{d}}{d}\right)\colon\mathcal{D}^{2}u_{j} u(𝐐+𝐈dd):𝒟2u\displaystyle\rightharpoonup u^{\star}\left(\mathbf{Q}^{\star}+\frac{\mathbf{I}_{d}}{d}\right)\colon\mathcal{D}^{2}u^{\star} in L1(Ω).\displaystyle\quad\text{in }L^{1}(\Omega).

Therefore, we have

lim infjΩ|𝒟2uj+q2(𝐐j+𝐈dd)uj|2\displaystyle\liminf_{j\to\infty}\int_{\Omega}\left|\mathcal{D}^{2}u_{j}+q^{2}\left(\mathbf{Q}_{j}+\frac{\mathbf{I}_{d}}{d}\right)u_{j}\right|^{2} =lim infjΩ(|𝒟2uj|2+2q2uj(𝐐j+𝐈dd):𝒟2uj+q4|uj(𝐐j+𝐈dd)|2)\displaystyle=\liminf_{j\to\infty}\int_{\Omega}\left(\left|\mathcal{D}^{2}u_{j}\right|^{2}+2q^{2}u_{j}\left(\mathbf{Q}_{j}+\frac{\mathbf{I}_{d}}{d}\right)\colon\mathcal{D}^{2}u_{j}+q^{4}\left|u_{j}\left(\mathbf{Q}_{j}+\frac{\mathbf{I}_{d}}{d}\right)\right|^{2}\right)
Ω(|𝒟2u|2+2q2u(𝐐+𝐈dd):𝒟2u+q4|u(𝐐+𝐈dd)|2)\displaystyle\geq\int_{\Omega}\left(\left|\mathcal{D}^{2}u^{\star}\right|^{2}+2q^{2}u^{\star}\left(\mathbf{Q}^{\star}+\frac{\mathbf{I}_{d}}{d}\right)\colon\mathcal{D}^{2}u^{\star}+q^{4}\left|u^{\star}\left(\mathbf{Q}^{\star}+\frac{\mathbf{I}_{d}}{d}\right)\right|^{2}\right)
=Ω|𝒟2u+q2(𝐐+𝐈dd)u|2.\displaystyle=\int_{\Omega}\left|\mathcal{D}^{2}u^{\star}+q^{2}\left(\mathbf{Q}^{\star}+\frac{\mathbf{I}_{d}}{d}\right)u^{\star}\right|^{2}. (8)

Hence, we can conclude that 𝒥(u,𝐐)\mathcal{J}(u^{\star},\mathbf{Q}^{\star}) achieves its minimum in the admissible space 𝒜\mathcal{A} by combining eq. 6, eq. 7 and eq. 8. ∎

Remark 1.2.

We briefly compare the proof with that of Bedford [4, Theorem 5.18]. In that work, the admissible space 𝒜BB\mathcal{A}^{BB} included a uniaxial constraint.This assumption is necessary to guarantee the 𝐇1\mathbf{H}^{1}-boundedness property of the minimising sequence {𝐐j}\{\mathbf{Q}_{j}\}, since that work seeks 𝐐SBV(Ω,S0)\mathbf{Q}\in SBV(\Omega,S_{0}) instead of 𝐇1\mathbf{H}^{1}. Enforcing this constraint numerically is difficult [6]; in this work we prefer the choice 𝐐𝐇1\mathbf{Q}\in\mathbf{H}^{1}, which enables us to remove the uniaxiality assumption.

2. A priori error estimates

We now consider the discretisation of the minimisation problem eq. 4. For simplicity, we analyse the decoupled case with q=0q=0, where eq. 4 splits into two independent problems: one for the smectic density variation uu:

minuH2Hb1(Ω,)𝒥1(u)=Ω(B|𝒟2u|2+fs(u)),\underset{u\in H^{2}\cap H^{1}_{b}(\Omega,\mathbb{R})}{\min}\quad\mathcal{J}_{1}(u)=\int_{\Omega}\left(B\left|\mathcal{D}^{2}u\right|^{2}+f_{s}(u)\right),

and one for the tensor field 𝐐\mathbf{Q},

min𝐐Hb1(Ω,S0)𝒥2(𝐐)=Ω(fn(𝐐,𝐐)).\underset{\mathbf{Q}\in H^{1}_{b}(\Omega,S_{0})}{\min}\quad\mathcal{J}_{2}(\mathbf{Q})=\int_{\Omega}\left(f_{n}(\mathbf{Q},\nabla\mathbf{Q})\right).

Here, Hb1(Ω,){uH1(Ω,):u=ub on Ω}H^{1}_{b}(\Omega,\mathbb{R})\coloneqq\{u\in H^{1}(\Omega,\mathbb{R}):u=u_{b}\text{ on }\partial\Omega\} and Hb1(Ω,S0){𝐐H1(Ω,S0):𝐐=𝐐b on Ω}H^{1}_{b}(\Omega,S_{0})\coloneqq\{\mathbf{Q}\in H^{1}(\Omega,S_{0}):\mathbf{Q}=\mathbf{Q}_{b}\text{ on }\partial\Omega\}. One can derive the following strong forms of their equilibrium equations using integration by parts. The optimality condition for uu yields a fourth-order partial differential equation (PDE)

{2B(𝒟2u)+a1u+a2u2+a3u3=0 in Ω,u=ub,𝒟2uν=𝒟2ubν on Ω,\begin{cases}2B\nabla\cdot\left(\nabla\cdot\mathcal{D}^{2}u\right)+a_{1}u+a_{2}u^{2}+a_{3}u^{3}=0&\text{ in }\Omega,\\ u=u_{b},\quad\mathcal{D}^{2}u\cdot\nu=\mathcal{D}^{2}u_{b}\cdot\nu&\text{ on }\partial\Omega,\end{cases}

where ν\nu denotes the outward unit normal. Note that the second boundary condition of uu is a natural boundary condition on the second derivative of uu. For simplicity, we only consider the case of a cubic nonlinearity (i.e., a2=0a_{2}=0) here as the quadratic term can be tackled similarly. Therefore, we consider the following governing equations

(𝒫1){2B(𝒟2u)+a1u+a3u3=0in Ω,u=ub,𝒟2uν=𝒟2ubνon Ω.(\mathcal{P}1)\quad\begin{cases}2B\nabla\cdot\left(\nabla\cdot\mathcal{D}^{2}u\right)+a_{1}u+a_{3}u^{3}=0&\text{in }\Omega,\\ u=u_{b},\quad\mathcal{D}^{2}u\cdot\nu=\mathcal{D}^{2}u_{b}\cdot\nu&\text{on }\partial\Omega.\end{cases} (9)

Meanwhile, the optimality condition for 𝐐\mathbf{Q} yields a second-order PDE

(𝒫2){d=2KΔ𝐐+2l(2|𝐐|21)𝐐=0 in Ω,d=3KΔ𝐐+l(𝐐|𝐐|2+2|𝐐|2𝐐)=0 in Ω,𝐐=𝐐b on Ω,(\mathcal{P}2)\quad\begin{cases}d=2\Rightarrow-K\Delta\mathbf{Q}+2l\left(2|\mathbf{Q}|^{2}-1\right)\mathbf{Q}=0&\text{ in }\Omega,\\ d=3\Rightarrow-K\Delta\mathbf{Q}+l\left(-\mathbf{Q}-|\mathbf{Q}|^{2}+2|\mathbf{Q}|^{2}\mathbf{Q}\right)=0&\text{ in }\Omega,\\ \mathbf{Q}=\mathbf{Q}_{b}&\text{ on }\partial\Omega,\end{cases} (10)

We now consider these two problems (𝒫1)(\mathcal{P}1) and (𝒫2)(\mathcal{P}2) in turn.

Remark 2.1.

The uniqueness of solutions is not expected. It is well-known that eq. 10 can support multiple solutions [25], while numerical experiments in [31] indicate the existence of multiple solutions to the optimality conditions for eq. 4.

2.1. A priori error estimates for (𝒫1)(\mathcal{P}1)

Since the PDE eq. 9 for the density variation uu is a fourth-order problem, a conforming discretisation requires a finite dimensional subspace of the Sobolev space H2(Ω)H^{2}(\Omega), which necessitates the use of 𝒞1\mathcal{C}^{1}-continuous elements. The construction of these elements is quite involved, particularly in three dimensions; without a special mesh structure, the lowest-degree conforming elements are the Argyris [1] and Zhang [32] elements, of degree 5 and 9 in two and three dimensions respectively. One approach to avoid such complexity is to use mixed formulations by solving two second-order systems, and we refer to [12, 27] for instance. However, this substantially increases the size of the linear systems to be solved. Alternatively, one can directly tackle the fourth-order problem with non-conforming elements, that do not satisfy the 𝒞1\mathcal{C}^{1}-requirement. For instance, the so-called continuous/discontinuous Galerkin methods and 𝒞0\mathcal{C}^{0} interior penalty methods (𝒞0\mathcal{C}^{0}-IP) are analysed in [9, 17], combining concepts from the theory of continuous and discontinuous Galerkin methods. Essentially, these methods use 𝒞0\mathcal{C}^{0}-conforming elements and penalise inter-element jumps in first derivatives to weakly enforce 𝒞1\mathcal{C}^{1}-continuity. This has the advantages of both convenience and efficiency: the weak form is simple, with only minor modifications from a conforming method, and fewer degrees of freedom are used than with a fully discontinuous Galerkin method.

We thus adopt the idea of 𝒞0\mathcal{C}^{0}-IP methods to solve the nonlinear fourth-order problem (𝒫1)(\mathcal{P}1) and derive a priori error estimates regarding uu. We adapt the techniques of [21] to prove convergence rates with the use of familiar continuous Lagrange elements for the problem (𝒫1)(\mathcal{P}1). The weak form of eq. 9 is defined as: find uH2(Ω)Hb1(Ω;)u\in H^{2}(\Omega)\cap H^{1}_{b}(\Omega;\mathbb{R}) such that

𝒩s(u)tAs(u,t)+Bs(u,u,u,t)+Cs(u,t)=Ls(t)tH2(Ω)H01(Ω),\mathcal{N}^{s}(u)t\coloneqq A^{s}(u,t)+B^{s}(u,u,u,t)+C^{s}(u,t)=L^{s}(t)\quad\forall t\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), (11)

where for t,wH2(Ω)t,w\in H^{2}(\Omega),

As(t,w)=2BΩ𝒟2t:𝒟2w,Cs(t,w)=a1Ωtw,Ls(t)2BΩ(𝒟2ubt)ν,A^{s}(t,w)=2B\int_{\Omega}\mathcal{D}^{2}t\colon\mathcal{D}^{2}w,\ C^{s}(t,w)=a_{1}\int_{\Omega}tw,\ L^{s}(t)\coloneqq 2B\int_{\partial\Omega}\left(\mathcal{D}^{2}u_{b}\cdot\nabla t\right)\cdot\nu,

and for μ,ζ,η,ξH2(Ω)\mu,\zeta,\eta,\xi\in H^{2}(\Omega),

Bs(μ,ζ,η,ξ)=a3Ωμζηξ.B^{s}(\mu,\zeta,\eta,\xi)=a_{3}\int_{\Omega}\mu\zeta\eta\xi.

Since eq. 11 is nonlinear, we derive its linearisation: find vH2(Ω)H01(Ω)v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega) such that

𝒟𝒩s(u)v,wH2As(v,w)+3Bs(u,u,v,w)+Cs(v,w)=Ls(w)wH2(Ω)H01(Ω),\langle\mathcal{D}\mathcal{N}^{s}(u)v,w\rangle_{H^{2}}\coloneqq A^{s}(v,w)+3B^{s}(u,u,v,w)+C^{s}(v,w)=L^{s}(w)\quad\forall w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), (12)

where ,H2\langle\cdot,\cdot\rangle_{H^{2}} represents the dual pairing between (H2(Ω)H01(Ω))\left(H^{2}(\Omega)\cap H^{1}_{0}(\Omega)\right)^{\star} and H2(Ω)H01(Ω)H^{2}(\Omega)\cap H^{1}_{0}(\Omega).

It is straightforward to derive the coercivity and boundedness of the bilinear operator As(,)A^{s}(\cdot,\cdot) with the semi-norm ||2|\cdot|_{2} (in fact, this is indeed a norm in H2(Ω)H01(Ω)H^{2}(\Omega)\cap H^{1}_{0}(\Omega)).

Lemma 2.2.

For v,wH2(Ω)H01(Ω)v,w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), there holds As(v,w)|v|2|w|2 and As(v,v)|v|22.A^{s}(v,w)\lesssim|v|_{2}|w|_{2}\text{ and }A^{s}(v,v)\gtrsim|v|_{2}^{2}.

Let 𝒯h\mathcal{T}_{h} be a mesh of Ω\Omega with TT denoting an element, and let I\mathcal{E}_{I} (resp. B\mathcal{E}_{B}) denote the set of all interior (resp. boundary) edges/faces ee of the mesh, and IB\mathcal{E}\coloneqq\mathcal{E}_{I}\cup\mathcal{E}_{B}. Define the broken Sobolev space H2(𝒯h){vH1(Ω):v|TH2(T)T𝒯h}H^{2}(\mathcal{T}_{h})\coloneqq\{v\in H^{1}(\Omega):v|_{T}\in H^{2}(T)\ \forall T\in\mathcal{T}_{h}\} equipped with the broken norm v2,𝒯h2=T𝒯hv2,T2\|v\|_{2,\mathcal{T}_{h}}^{2}=\sum_{T\in\mathcal{T}_{h}}\|v\|^{2}_{2,T}. We take a nonconforming but continuous approximation uhu_{h} for the solution uu of eq. 11, that is to say, uhWh,bH2(𝒯h)Hb1(Ω)u_{h}\in W_{h,b}\subset H^{2}(\mathcal{T}_{h})\cap H^{1}_{b}(\Omega) defined for deg2\deg\geq 2 (since (𝒫1)(\mathcal{P}1) is a fourth-order problem) by

Wh\displaystyle W_{h} {vH2(𝒯h)H1(Ω):vdeg(T)T𝒯h},\displaystyle\coloneqq\{v\in H^{2}(\mathcal{T}_{h})\cap H^{1}(\Omega):v\in\mathbb{Q}_{\deg}(T)\ \forall T\in\mathcal{T}_{h}\},
Wh,0\displaystyle W_{h,0} {vH2(𝒯h)H1(Ω):v=0 on Ω,vdeg(T)T𝒯h},\displaystyle\coloneqq\{v\in H^{2}(\mathcal{T}_{h})\cap H^{1}(\Omega):v=0\text{ on }\partial\Omega,v\in\mathbb{Q}_{\deg}(T)\ \forall T\in\mathcal{T}_{h}\},
Wh,b\displaystyle W_{h,b} {vH2(𝒯h)H1(Ω):v=ub on Ω,vdeg(T)T𝒯h},\displaystyle\coloneqq\{v\in H^{2}(\mathcal{T}_{h})\cap H^{1}(\Omega):v=u_{b}\text{ on }\partial\Omega,v\in\mathbb{Q}_{\deg}(T)\ \forall T\in\mathcal{T}_{h}\},

with deg\mathbb{Q}_{\deg} denoting piecewise continuous polynomials of degree deg{\deg} on a mesh of quadrilaterals or hexahedra. Following the derivation of the 𝒞0\mathcal{C}^{0}-IP formulation given in [7, Section 3], we introduce the discrete nonlinear weak form: find uhWh,bu_{h}\in W_{h,b} such that

𝒩hs(uh)thAhs(uh,th)+Phs(uh,th)+Bs(uh,uh,uh,th)+Cs(uh,th)=Ls(th)thWh,0,\mathcal{N}_{h}^{s}(u_{h})t_{h}\coloneqq A^{s}_{h}(u_{h},t_{h})+P^{s}_{h}(u_{h},t_{h})+B^{s}(u_{h},u_{h},u_{h},t_{h})+C^{s}(u_{h},t_{h})=L^{s}(t_{h})\quad\forall t_{h}\in W_{h,0}, (13)

where for all uh,thWhu_{h},t_{h}\in W_{h},

Ahs(uh,th)2B(T𝒯hT𝒟2uh:𝒟2theIe{{2uhν2}}theIe{{2thν2}}uh),A^{s}_{h}(u_{h},t_{h})\coloneqq 2B\bigg{(}\sum_{T\in\mathcal{T}_{h}}\int_{T}\mathcal{D}^{2}u_{h}\colon\mathcal{D}^{2}t_{h}-\sum_{e\in\mathcal{E}_{I}}\int_{e}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}u_{h}}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla t_{h}\rrbracket-\sum_{e\in\mathcal{E}_{I}}\int_{e}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}t_{h}}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla u_{h}\rrbracket\bigg{)},

and

Phs(uh,th)eI2Bϵhe3euhth.P^{s}_{h}(u_{h},t_{h})\coloneqq\sum_{e\in\mathcal{E}_{I}}\frac{2B\epsilon}{h_{e}^{3}}\int_{e}\llbracket\nabla u_{h}\rrbracket\llbracket\nabla t_{h}\rrbracket. (14)

Here, ϵ\epsilon is the penalty parameter (to be specified in section 3), heh_{e} indicates the size of the edge/face ee and the average of the second derivatives of uu along the normal direction across the edge/facet ee is defined as {{2uν2}}=12(2u+ν2|e+2uν2|e).\left\{\mskip-5.0mu\left\{\frac{\partial^{2}u}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}=\frac{1}{2}\left(\left.\frac{\partial^{2}u_{+}}{\partial\nu^{2}}\right|_{e}+\left.\frac{\partial^{2}u_{-}}{\partial\nu^{2}}\right|_{e}\right). For any interior edge eIe\in\mathcal{E}_{I} shared by two cells TT_{-} and T+T_{+}, we define the jump 𝐯\llbracket\mathbf{v}\rrbracket by 𝐯=𝐯ν+𝐯+ν+\llbracket\mathbf{v}\rrbracket=\mathbf{v}_{-}\cdot\nu_{-}+\mathbf{v}_{+}\cdot\nu_{+} with ν,ν+\nu_{-},\nu_{+} representing the restriction of outward normals in T,T+T_{-},T_{+} respectively. On the boundary edge/face eBe\in\mathcal{E}_{B}, we define 𝐯=𝐯ν\llbracket\mathbf{v}\rrbracket=\mathbf{v}\cdot\nu. The operator PhsP^{s}_{h} penalises the first derivatives across the interior edge/facet since the function in H1(Ω)H^{1}(\Omega) is not necessarily continuously differentiable.

The linearised version is to seek vhWh,0v_{h}\in W_{h,0} such that

𝒟𝒩hs(uh)vh,wh=Ls(wh)whWh,0,\langle\mathcal{D}\mathcal{N}_{h}^{s}(u_{h})v_{h},w_{h}\rangle=L^{s}(w_{h})\quad\forall w_{h}\in W_{h,0}, (15)

where

𝒟𝒩hs(uh)vh,whAhs(vh,wh)+Phs(vh,wh)+3Bs(uh,uh,vh,wh)+Cs(vh,wh).\langle\mathcal{D}\mathcal{N}_{h}^{s}(u_{h})v_{h},w_{h}\rangle\coloneqq A_{h}^{s}(v_{h},w_{h})+P_{h}^{s}(v_{h},w_{h})+3B^{s}(u_{h},u_{h},v_{h},w_{h})+C^{s}(v_{h},w_{h}). (16)

We also define the mesh-dependent H2H^{2}-like semi-norm for vWhv\in W_{h},

|||v|||h2T𝒯h|v|H2(T)2+eIe1he3|v|2.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\coloneqq\sum_{T\in\mathcal{T}_{h}}|v|_{H^{2}(T)}^{2}+\sum_{e\in\mathcal{E}_{I}}\int_{e}\frac{1}{h_{e}^{3}}|\llbracket\nabla v\rrbracket|^{2}. (17)

Note that ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} is indeed a norm on Wh,0W_{h,0}. This norm will be used in the well-posedness and convergence analysis below.

We first give an immediate result about the consistency of the discrete form eq. 13.

Theorem 2.3.

(Consistency) Assuming that uH4(Ω)u\in H^{4}(\Omega). The solution uu of the continuous weak form eq. 11 solves the discrete weak problem eq. 13.

Proof.

Multiplying the fourth-order term 2B((𝒟2u))2B\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}u)) in eq. 9 with a test function tWh,0t\in W_{h,0} and using piecewise integration by parts with the boundary condition specified in eq. 9 for uu, one can obtain

2BThT((𝒟2u))t=2BThT𝒟2u:𝒟2t2BeIe{{2uν2}}t.2B\sum_{T\in\mathcal{E}_{h}}\int_{T}\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}u))t=2B\sum_{T\in\mathcal{E}_{h}}\int_{T}\mathcal{D}^{2}u\colon\mathcal{D}^{2}t-2B\sum_{e\in\mathcal{E}_{I}}\int_{e}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}u}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla t\rrbracket. (18)

Since uH4(Ω)u\in H^{4}(\Omega) implies u\nabla u is continuous on the whole domain Ω\Omega, the jump term u\llbracket\nabla u\rrbracket then becomes zero and we can thus symmetrise and penalise the form eq. 18. This leads to the presence of Ahs(u,t)+Phs(u,t)A^{s}_{h}(u,t)+P^{s}_{h}(u,t). The remaining terms involving BsB^{s} and CsC^{s} are straightforward as one takes the test function tWh,0t\in W_{h,0}. Therefore, uu satisfies eq. 13. ∎

By noting that (𝒟2:𝒟2)u=[(x2)2+(y2)2+2(xy2)2]u=Δ2u,\left(\mathcal{D}^{2}\colon\mathcal{D}^{2}\right)u=\left[\left(\partial_{x}^{2}\right)^{2}+\left(\partial_{y}^{2}\right)^{2}+2\left(\partial_{xy}^{2}\right)^{2}\right]u=\Delta^{2}u, it is natural to extend the classical elliptic regularity result [5] for the biharmonic operator Δ2\Delta^{2} to the case of the bi-Hessian operator 𝒟2:𝒟2\mathcal{D}^{2}\colon\mathcal{D}^{2}. If the domain Ω\Omega is smooth, the weak solutions of the biharmonic problem (e.g., [7, Example 2]) belong to H4(Ω)H^{4}(\Omega) by classical elliptic regularity results and thus we make this assumption henceforth.

Moreover, to facilitate the analysis, we further assume that uu is an isolated solution, i.e., there is only one solution uu satisfying eq. 9 within a sufficiently small ball {vH2(Ω)H01(Ω):|vu|2rb}\{v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega):|v-u|_{2}\leq r_{b}\} with radius rbr_{b}. These assumptions then imply that the linearised operator 𝒟𝒩s(u),H2\langle\mathcal{D}\mathcal{N}^{s}(u)\cdot,\cdot\rangle_{H^{2}} satisfies the following inf-sup condition:

0<βu=infvH2(Ω)H01(Ω)|v|2=1supwH2(Ω)H01(Ω)|w|2=1𝒟𝒩s(u)v,wH2=infwH2(Ω)H01(Ω)|w|2=1supvH2(Ω)H01(Ω)|v|2=1𝒟𝒩s(u)v,wH2.0<\beta_{u}=\adjustlimits{\inf}_{\underset{|v|_{2}=1}{v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega)}}{\sup}_{\underset{|w|_{2}=1}{w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega)}}\langle\mathcal{D}\mathcal{N}^{s}(u)v,w\rangle_{H^{2}}=\adjustlimits{\inf}_{\underset{|w|_{2}=1}{w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega)}}{\sup}_{\underset{|v|_{2}=1}{v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega)}}\langle\mathcal{D}\mathcal{N}^{s}(u)v,w\rangle_{H^{2}}. (19)

2.1.1. Well-posedness of the discrete form

Recalling [7, Eq. (3.20)], we can obtain for v,wWh,0v,w\in W_{h,0},

eI|e{{2wν2}}v|(T𝒯hT𝒟2w:𝒟2w)1/2(eI1he3e(v)2)1/2,\sum_{e\in\mathcal{E}_{I}}\left|\int_{e}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}w}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla v\rrbracket\right|\lesssim\left(\sum_{T\in\mathcal{T}_{h}}\int_{T}\mathcal{D}^{2}w\colon\mathcal{D}^{2}w\right)^{1/2}\left(\sum_{e\in\mathcal{E}_{I}}\frac{1}{h_{e}^{3}}\int_{e}(\llbracket\nabla v\rrbracket)^{2}\right)^{1/2}, (20)

as the edge/facet size he<1h_{e}<1. With the estimate eq. 20 at hand, we then apply the Cauchy–Schwarz inequality and use the definition eq. 17 of ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} to obtain the boundedness of Ahs(,)A^{s}_{h}(\cdot,\cdot) and Phs(,)P_{h}^{s}(\cdot,\cdot). We omit the details of the proofs here and only illustrate the boundedness result for Bs(,,,)B^{s}(\cdot,\cdot,\cdot,\cdot) and Cs(,)C^{s}(\cdot,\cdot) below.

Lemma 2.4.

(Boundedness of Bs(,,,)B^{s}(\cdot,\cdot,\cdot,\cdot) and Cs(,)C^{s}(\cdot,\cdot).) For u,v,w,pWh,0u,v,w,p\in W_{h,0}, we have

|Bs(u,v,w,p)||u|h|v|h|w|h|p|h and |Cs(u,v)||u|h|v|h.|B^{s}(u,v,w,p)|\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|p\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\text{ and }|C^{s}(u,v)|\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. (21)

For u,vH2(Ω)u,v\in H^{2}(\Omega), w,pWhw,p\in W_{h},

|Bs(u,v,w,p)|u2v2|w|h|p|h.|B^{s}(u,v,w,p)|\lesssim\|u\|_{2}\|v\|_{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|p\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. (22)
Proof.

By Hölder’s inequality, the Sobolev embedding H1(Ω)L4(Ω)H^{1}(\Omega)\hookrightarrow L^{4}(\Omega), and the fact that the H1H^{1} semi-norm ||1|\cdot|_{1} is a norm in H01(Ω)H^{1}_{0}(\Omega), we deduce

|Bs(u,v,w,p)|\displaystyle|B^{s}(u,v,w,p)| uL4vL4wL4pL4\displaystyle\lesssim\|u\|_{L^{4}}\|v\|_{L^{4}}\|w\|_{L^{4}}\|p\|_{L^{4}}
|u|1|v|1|w|1|p|1.\displaystyle\lesssim|u|_{1}|v|_{1}|w|_{1}|p|_{1}.

It then follows from a Poincaré inequality [11, Eq. (5.7)] for piecewise H2H^{2} functions that

T𝒯h|v|1,T2T𝒯h|v|2,T2+eI1he3v0,e2=|||v|||h2vWh,0.\sum_{T\in\mathcal{T}_{h}}|v|^{2}_{1,T}\lesssim\sum_{T\in\mathcal{T}_{h}}|v|^{2}_{2,T}+\sum_{e\in\mathcal{E}_{I}}\frac{1}{h_{e}^{3}}\|\llbracket\nabla v\rrbracket\|^{2}_{0,e}={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}_{h}\quad\forall v\in W_{h,0}. (23)

Thus, we obtain |Bs(u,v,w,p)||u|h|v|h|w|h|p|h.|B^{s}(u,v,w,p)|\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|p\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

The boundedness of Cs(,)C^{s}(\cdot,\cdot) follows similarly by the Cauchy–Schwarz inequality, the Sobolev embedding H1(Ω)L2(Ω)H^{1}(\Omega)\hookrightarrow L^{2}(\Omega) and the use of eq. 23. The proof of eq. 22 is analogous to that of eq. 21 with a use of the embedding result H2(Ω)L(Ω)H^{2}(\Omega)\hookrightarrow L^{\infty}(\Omega) and the Cauchy–Schwarz inequality. ∎

We give the coercivity result for the bilinear form (Ahs(,)+Phs(,))\left(A^{s}_{h}(\cdot,\cdot)+P^{s}_{h}(\cdot,\cdot)\right).

Lemma 2.5.

(Coercivity of Ahs+PhsA^{s}_{h}+P^{s}_{h}) For a sufficiently large penalty parameter ϵ\epsilon, there holds

|vh|h2Ahs(vh,vh)+Phs(vh,vh)vhWh,0.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\lesssim A^{s}_{h}(v_{h},v_{h})+P^{s}_{h}(v_{h},v_{h})\quad\forall v_{h}\in W_{h,0}. (24)
Proof.

By eq. 20 and the inequality of geometric and arithmetic means, we deduce for vWhv\in W_{h},

Ahs(v,v)+Phs(v,v)\displaystyle A^{s}_{h}(v,v)+P^{s}_{h}(v,v) 2BT𝒯h|v|H2(T)22BC(T𝒯h|v|2,T2)1/2(eI1he3v0,e2)1/2+2B(eIeϵhe3|v|2)\displaystyle\geq 2B\sum_{T\in\mathcal{T}_{h}}|v|_{H^{2}(T)}^{2}-2BC\left(\sum_{T\in\mathcal{T}_{h}}|v|^{2}_{2,T}\right)^{1/2}\left(\sum_{e\in\mathcal{E}_{I}}\frac{1}{h_{e}^{3}}\|\llbracket\nabla v\rrbracket\|^{2}_{0,e}\right)^{1/2}+2B\left(\sum_{e\in\mathcal{E}_{I}}\int_{e}\frac{\epsilon}{h_{e}^{3}}|\llbracket\nabla v\rrbracket|^{2}\right)
2B[12T𝒯h|v|H2(T)2+(ϵC22)eI1he3v0,e2]B|||v|||h2,\displaystyle\geq 2B\left[\frac{1}{2}\sum_{T\in\mathcal{T}_{h}}|v|_{H^{2}(T)}^{2}+\left(\epsilon-\frac{C^{2}}{2}\right)\sum_{e\in\mathcal{E}_{I}}\frac{1}{h_{e}^{3}}\|\llbracket\nabla v\rrbracket\|^{2}_{0,e}\right]\geq B{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2},

provided the penalty parameter ϵ\epsilon is sufficiently large with the generic constant CC from eq. 20. ∎

An important question about the well-posedness is the coercivity of the bilinear operator 𝒟𝒩hs(uh),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u_{h})\cdot,\cdot\rangle. Due to the presence of BsB^{s} and CsC^{s} terms in 𝒟𝒩hs(uh),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u_{h})\cdot,\cdot\rangle, it is not trivial to derive its coercivity. We first discuss the weak coercivity of the bilinear form 𝒟𝒩hs(u),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\cdot,\cdot\rangle defined as

𝒟𝒩hs(u)vh,whAhs(vh,wh)+Phs(vh,wh)+3Bs(u,u,vh,wh)+Cs(vh,wh)vh,whWh,0.\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle\coloneqq A^{s}_{h}(v_{h},w_{h})+P^{s}_{h}(v_{h},w_{h})+3B^{s}(u,u,v_{h},w_{h})+C^{s}(v_{h},w_{h})\quad\forall v_{h},w_{h}\in W_{h,0}. (25)

To this end, we will employ the enrichment operator Eh:WhWCH2(Ω)E_{h}:W_{h}\rightarrow W_{C}\subset H^{2}(\Omega) with WCW_{C} being the Hsieh–Clough–Tocher macro finite element space [7]. The following lemma is adapted to our notation and definition of ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} from [7, Lemma 1].

Lemma 2.6.

[7, Lemma 1] For vhWh,0v_{h}\in W_{h,0}, there holds

T𝒯h(h4vhEhvhL2(T)2+h2|vhEhvh|H1(T)2+|vhEhvh|H2(T)2)eI1he3vhL2(e)2|||vh|||h2.\sum_{T\in\mathcal{T}_{h}}\left(h^{-4}\|v_{h}-E_{h}v_{h}\|^{2}_{L^{2}(T)}+h^{-2}|v_{h}-E_{h}v_{h}|^{2}_{H^{1}(T)}+|v_{h}-E_{h}v_{h}|^{2}_{H^{2}(T)}\right)\lesssim\sum_{e\in\mathcal{E}_{I}}\frac{1}{h_{e}^{3}}\|\llbracket\nabla v_{h}\rrbracket\|^{2}_{L^{2}(e)}\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}.

With this, we obtain the following discrete inf-sup condition for the discrete bilinear operator 𝒟𝒩hs(u),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\cdot,\cdot\rangle.

Theorem 2.7.

(Weak coercivity of 𝒟𝒩hs(u),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\cdot,\cdot\rangle) Let uu be a regular isolated solution of the nonlinear continuous weak form eq. 13. For a sufficiently large ϵ\epsilon and a sufficiently small mesh size hh, the following discrete inf-sup condition holds with a positive constant βc>0\beta_{c}>0:

0<βcinfvWh,0|vh|h=1supwWh,0|wh|h=1𝒟𝒩hs(u)vh,wh.0<\beta_{c}\leq\adjustlimits{\inf}_{\underset{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|v_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}{v\in W_{h,0}}}{\sup}_{\underset{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|w_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}{w\in W_{h,0}}}\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle. (26)
Proof.

For vH2(Ω)H01(Ω)v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), it follows from the boundedness result of Bs,CsB^{s},C^{s} that Bs(u,u,v,)B^{s}(u,u,v,\cdot) and Cs(v,)(L2(Ω))C^{s}(v,\cdot)\in(L^{2}(\Omega))^{\star}. Furthermore, since As(,)A^{s}(\cdot,\cdot) is bounded and coercive as given by lemma 2.2, for a given vhWhv_{h}\in W_{h} with |vh|h=1{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}=1, there exists ξ\xi and ηH4(Ω)H01(Ω)\eta\in H^{4}(\Omega)\cap H^{1}_{0}(\Omega) that solve the linear systems:

As(ξ,w)=3Bs(u,u,vh,w)+Cs(vh,w)wH2(Ω)H01(Ω),A^{s}(\xi,w)=3B^{s}(u,u,v_{h},w)+C^{s}(v_{h},w)\quad\forall w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), (27a)
As(η,w)=3Bs(u,u,Ehvh,w)+Cs(Ehvh,w)wH2(Ω)H01(Ω).A^{s}(\eta,w)=3B^{s}(u,u,E_{h}v_{h},w)+C^{s}(E_{h}v_{h},w)\quad\forall w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega). (27b)

We require the H4H^{4}-regularity for the derivation of eq. 34 below. It then follows from the standard elliptic regularity result that η4CBC\|\eta\|_{4}\lesssim C_{BC} with a constant CBCC_{BC} depending on u2\|u\|_{2}.

Subtracting eq. 27a from eq. 27b, then taking w=ηξw=\eta-\xi and using the coercivity of As(,)A^{s}(\cdot,\cdot), boundedness of Bs(,,,)B^{s}(\cdot,\cdot,\cdot,\cdot), Cs(,)C^{s}(\cdot,\cdot) and enrichment estimate lemma 2.6, we get

|ηξ|2(3u22+1)Ehvhvh0h2|vh|h=h2.|\eta-\xi|_{2}\lesssim\left(3\|u\|_{2}^{2}+1\right)\|E_{h}v_{h}-v_{h}\|_{0}\lesssim h^{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}=h^{2}. (28)

Since uu is a regular isolated solution of eq. 11, it yields by eq. 19, eq. 27b, lemma 2.2, the fact that (Ehvh+η)=0\llbracket\nabla(E_{h}v_{h}+\eta)\rrbracket=0 and the triangle inequality, that there exists wH2(Ω)H01(Ω)w\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega) with |w|2=1|w|_{2}=1 such that

|Ehvh|2\displaystyle|E_{h}v_{h}|_{2} 𝒟𝒩s(u)Ehvh,wH2=As(Ehvh,w)+3Bs(u,u,Ehvh,w)+Cs(Ehvh,w)\displaystyle\lesssim\langle\mathcal{D}\mathcal{N}^{s}(u)E_{h}v_{h},w\rangle_{H^{2}}=A^{s}(E_{h}v_{h},w)+3B^{s}(u,u,E_{h}v_{h},w)+C^{s}(E_{h}v_{h},w)
=As(Ehvh+η,w)|Ehvh+η|2|w|2=|Ehvh+η|h\displaystyle=A^{s}(E_{h}v_{h}+\eta,w)\lesssim|E_{h}v_{h}+\eta|_{2}{|w|_{2}}={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}+\eta\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}
|Ehvhvh|h+|vh+Ihξ|h+|Ihξξ|h+|ξη|h=|ξη|2.\displaystyle\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}-v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+I_{h}\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}\xi-\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+\underbrace{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\xi-\eta\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}_{=|\xi-\eta|_{2}}. (29)

Note that ξ=0\llbracket\nabla\xi\rrbracket=0 on I\mathcal{E}_{I} since ξH4(Ω)\xi\in H^{4}(\Omega). We can thus calculate using lemma 2.6 that

|||Ehvhvh|||h2eIe1he3|vh|2eIe1he3|(vh+ξ)|2|||vh+ξ|||h2.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}-v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\lesssim\sum_{e\in\mathcal{E}_{I}}\int_{e}\frac{1}{h_{e}^{3}}|\llbracket\nabla v_{h}\rrbracket|^{2}\lesssim\sum_{e\in\mathcal{E}_{I}}\int_{e}\frac{1}{h_{e}^{3}}|\llbracket\nabla(v_{h}+\xi)\rrbracket|^{2}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}.

Further, by the triangle inequality, we get

|Ehvhvh|h|vh+ξ|h|vh+Ihξ|h+|ξIhξ|h.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}-v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+I_{h}\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\xi-I_{h}\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. (30)

Since vh+IhξWhv_{h}+I_{h}\xi\in W_{h}, it follows from the coercivity result eq. 24 that there exists whWhw_{h}\in W_{h} with |w|h=1{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}=1 such that

|vh+Ihξ|h\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+I_{h}\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} Ahs(vh+Ihξ,wh)+Phs(vh+Ihξ,wh)\displaystyle\lesssim A^{s}_{h}(v_{h}+I_{h}\xi,w_{h})+P^{s}_{h}(v_{h}+I_{h}\xi,w_{h})
=𝒟𝒩hs(u)vh,wh3Bs(u,u,vh,wh)Cs(vh,wh)\displaystyle=\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle-3B^{s}(u,u,v_{h},w_{h})-C^{s}(v_{h},w_{h})
+Ahs(Ihξξ,wh)+Phs(Ihξξ,wh)+Ahs(ξ,wh)+Phs(ξ,wh)\displaystyle\quad+A^{s}_{h}(I_{h}\xi-\xi,w_{h})+P^{s}_{h}(I_{h}\xi-\xi,w_{h})+A^{s}_{h}(\xi,w_{h})+P^{s}_{h}(\xi,w_{h})
=𝒟𝒩hs(u)vh,wh+3Bs(u,u,vh,Ehwhwh)+Cs(vh,Ehwhwh)\displaystyle=\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+3B^{s}(u,u,v_{h},E_{h}w_{h}-w_{h})+C^{s}(v_{h},E_{h}w_{h}-w_{h})
+Ahs(Ihξξ,wh)+Phs(Ihξξ,wh)+Ahs(ξ,whEhwh)+Phs(ξ,whEhwh),\displaystyle\quad+A^{s}_{h}(I_{h}\xi-\xi,w_{h})+P^{s}_{h}(I_{h}\xi-\xi,w_{h})+A^{s}_{h}(\xi,w_{h}-E_{h}w_{h})+P^{s}_{h}(\xi,w_{h}-E_{h}w_{h}), (31)

where in the last equality we have used the fact that

3Bs(u,u,vh,Ehwh)+Cs(vh,Ehwh)=As(ξ,Ehwh)=Ahs(ξ,Ehwh)+Ps(ξ,Ehwh)3B^{s}(u,u,v_{h},E_{h}w_{h})+C^{s}(v_{h},E_{h}w_{h})=A^{s}(\xi,E_{h}w_{h})=A^{s}_{h}(\xi,E_{h}w_{h})+P^{s}(\xi,E_{h}w_{h})

because of eq. 27a and ξ=Ehwh=0\llbracket\nabla\xi\rrbracket=\llbracket\nabla E_{h}w_{h}\rrbracket=0.

Using the boundedness result lemma 2.4 and the enrichment estimate lemma 2.6, we obtain

3Bs(u,u,vh,Ehwhwh)+Cs(vh,Ehwhwh)vh0|vh|1|vh|h=1Ehwhwh0h2|wh|h=h2.3B^{s}(u,u,v_{h},E_{h}w_{h}-w_{h})+C^{s}(v_{h},E_{h}w_{h}-w_{h})\lesssim\underbrace{\|v_{h}\|_{0}}_{\lesssim|v_{h}|_{1}\lesssim{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|v_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}\underbrace{\|E_{h}w_{h}-w_{h}\|_{0}}_{\lesssim h^{2}{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|w_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=h^{2}}. (32)

By the boundedness of the bilinear form Ahs+PhsA^{s}_{h}+P^{s}_{h} and standard interpolation estimate, we have

Ahs(Ihξξ,wh)+Phs(Ihξξ,wh)|Ihξξ|h|wh|h=1hmin{deg1,2}ξ4.A^{s}_{h}(I_{h}\xi-\xi,w_{h})+P^{s}_{h}(I_{h}\xi-\xi,w_{h})\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}\xi-\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\underbrace{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}_{=1}\lesssim h^{\min\{\deg-1,2\}}\|\xi\|_{4}. (33)

Moreover, by the fact that ξ=(Ehwh)=0\llbracket\nabla\xi\rrbracket=\llbracket\nabla(E_{h}w_{h})\rrbracket=0, the enrichment estimate lemma 2.6 and eq. 18, there holds

Ahs\displaystyle A^{s}_{h} (ξ,whEhwh)+Phs(ξ,whEhwh)\displaystyle(\xi,w_{h}-E_{h}w_{h})+P^{s}_{h}(\xi,w_{h}-E_{h}w_{h})
=2BT𝒯hT𝒟2ξ:𝒟2(whEhwh)2BeIe{{2ξν2}}(whEhwh)\displaystyle=2B\sum_{T\in\mathcal{T}_{h}}\int_{T}\mathcal{D}^{2}\xi\colon\mathcal{D}^{2}(w_{h}-E_{h}w_{h})-2B\sum_{e\in\mathcal{E}_{I}}\int_{e}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}\xi}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla(w_{h}-E_{h}w_{h})\rrbracket
=2BT𝒯h((𝒟2ξ))(whEhwh)ξ4whEhwh0h2ξ4.\displaystyle=2B\sum_{T\in\mathcal{T}_{h}}\nabla\cdot\left(\nabla\cdot(\mathcal{D}^{2}\xi)\right)(w_{h}-E_{h}w_{h})\lesssim\|\xi\|_{4}\|w_{h}-E_{h}w_{h}\|_{0}\lesssim h^{2}\|\xi\|_{4}. (34)

Combine eqs. 32, 33 and 34 in eq. 31 to obtain

|vh+Ihξ|h𝒟𝒩hs(u)vh,wh+h2+hmin{deg1,2}.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}+I_{h}\xi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+h^{2}+h^{\min\{\deg-1,2\}}. (35)

Substituting eq. 35 into eq. 30 and using standard interpolation estimates yield that

|Ehvhvh|h𝒟𝒩hs(u)vh,wh+h2+hmin{deg1,2}.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}-v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+h^{2}+h^{\min\{\deg-1,2\}}. (36)

A use of eqs. 36 and 35, standard interpolation estimates and eq. 28 in eq. 29 leads to

|Ehvh|2𝒟𝒩hs(u)vh,wh+h2+hmin{deg1,2}.|E_{h}v_{h}|_{2}\lesssim\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+h^{2}+h^{\min\{\deg-1,2\}}.

Then, by the triangle inequality, we have

1=|vh|h|vhEhvh|h+|Ehvh|h=|Ehvh|2Ct(𝒟𝒩hs(u)vh,wh+h2+hmin{deg1,2}).1={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}-E_{h}v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+\underbrace{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|E_{h}v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}_{=|E_{h}v_{h}|_{2}}\leq C_{t}\left(\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+h^{2}+h^{\min\{\deg-1,2\}}\right).

Therefore, for the mesh size hh satisfying h2+hmin{deg1,2}<12Ct,h^{2}+h^{\min\{\deg-1,2\}}<\frac{1}{2C_{t}}, the discrete inf-sup condition eq. 26 holds for βc=12Ct\beta_{c}=\frac{1}{2C_{t}}. ∎

We can now obtain the discrete inf-sup condition for the perturbed bilinear form 𝒟𝒩hs(Ihu),\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\cdot,\cdot\rangle given by

𝒟𝒩hs(Ihu)vh,wh=Ahs(vh,wh)+Phs(vh,wh)+3Bs(Ihu,Ihu,vh,wh)+Cs(vh,wh)vh,whWh,0.\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)v_{h},w_{h}\rangle=A^{s}_{h}(v_{h},w_{h})+P^{s}_{h}(v_{h},w_{h})+3B^{s}(I_{h}u,I_{h}u,v_{h},w_{h})+C^{s}(v_{h},w_{h})\quad\forall v_{h},w_{h}\in W_{h,0}. (37)

Here, we employ the interpolation operator Ih:H2(Ω)Hb1(Ω;)Wh,bI_{h}:H^{2}(\Omega)\cap H^{1}_{b}(\Omega;\mathbb{R})\to W_{h,b}.

Theorem 2.8.

(Weak coercivity of 𝒟𝒩hs(Ihu),\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\cdot,\cdot\rangle) Let uu be a regular isolated solution of the nonlinear continuous weak form eq. 13 and IhuI_{h}u the interpolation of uu. For a sufficiently large ϵ\epsilon and a sufficiently small mesh size hh, the following discrete inf-sup condition holds:

0<βc2infvhWh,0|vh|h=1supwhWh,0|wh|h=1𝒟𝒩hs(Ihu)vh,wh.0<\frac{\beta_{c}}{2}\leq\adjustlimits{\inf}_{\underset{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|v_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}{v_{h}\in W_{h,0}}}{\sup}_{\underset{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|w_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}{w_{h}\in W_{h,0}}}\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)v_{h},w_{h}\rangle. (38)
Proof.

Denote u~=uIhu\tilde{u}=u-I_{h}u. By the definition eq. 37 of the bilinear form 𝒟𝒩hs(Ihu),\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\cdot,\cdot\rangle, we have 𝒟𝒩hs(Ihu)vh,wh=Ahs(vh,wh)+Phs(vh,wh)+3Bs(uu~,uu~,vh,wh)+Cs(vh,wh).\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)v_{h},w_{h}\rangle=A^{s}_{h}(v_{h},w_{h})+P^{s}_{h}(v_{h},w_{h})+3B^{s}(u-\tilde{u},u-\tilde{u},v_{h},w_{h})+C^{s}(v_{h},w_{h}). It follows from the definition of BsB^{s} and its boundedness result lemma 2.4 that

Bs(uu~,uu~,vh,wh)\displaystyle B^{s}(u-\tilde{u},u-\tilde{u},v_{h},w_{h}) =Bs(u,u,vh,wh)+Bs(u~,u~,vh,wh)2Bs(u,u~,vh,wh)\displaystyle=B^{s}(u,u,v_{h},w_{h})+B^{s}(\tilde{u},\tilde{u},v_{h},w_{h})-2B^{s}(u,\tilde{u},v_{h},w_{h})
Bs(u,u,vh,wh)+Bs(u~,u~,vh,wh)2C1|u|h|u~|h|vh|h|wh|h,\displaystyle\geq B^{s}(u,u,v_{h},w_{h})+B^{s}(\tilde{u},\tilde{u},v_{h},w_{h})-2C_{1}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\tilde{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h},

where C1C_{1} is the generic constant arising in the boundedness result lemma 2.4 for Bs(,,,)B^{s}(\cdot,\cdot,\cdot,\cdot). Therefore, we obtain that

𝒟𝒩hs(Ihu)vh,wh𝒟𝒩hs(u)vh,wh+3Bs(u~,u~,vh,wh)6C1|u|h|u~|h|vh|h|wh|h.\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)v_{h},w_{h}\rangle\geq\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle+3B^{s}(\tilde{u},\tilde{u},v_{h},w_{h})-6C_{1}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\tilde{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

Now using the inf-sup condition theorem 2.7 for the bilinear form 𝒟𝒩hs(u),\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\cdot,\cdot\rangle, boundedness result lemma 2.4 and interpolation estimates, we get

sup|wh|h=1whWh,0𝒟𝒩hs(Ihu)vh,wh\displaystyle\sup_{\underset{w_{h}\in W_{h,0}}{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|w_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}}\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)v_{h},w_{h}\rangle sup|wh|h=1whWh,0𝒟𝒩hs(u)vh,wh\displaystyle\geq\sup_{\underset{w_{h}\in W_{h,0}}{{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|w_{h}\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}_{h}=1}}\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)v_{h},w_{h}\rangle
3|Bs(u~,u~,vh,wh)|6C1hmin{deg1,𝕜u2}|u|h|vh|h\displaystyle\quad-3|B^{s}(\tilde{u},\tilde{u},v_{h},w_{h})|-6C_{1}h^{\min\{\deg-1,\Bbbk_{u}-2\}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}
(βcC2hmin{deg1,𝕜u2})|vh|hβc2|vh|h,\displaystyle\geq\left(\beta_{c}-C_{2}h^{\min\{\deg-1,\Bbbk_{u}-2\}}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\geq\frac{\beta_{c}}{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h},

for a sufficiently small mesh size hh such that hmin{deg1,𝕜u2}<βc2C2h^{\min\{\deg-1,\Bbbk_{u}-2\}}<\frac{\beta_{c}}{2C_{2}}. Here, C2C_{2} depends on C1C_{1} and u𝕜u\|u\|_{\Bbbk_{u}} and 𝕜u4\Bbbk_{u}\geq 4 gives the regularity of uu, i.e., uH𝕜u(Ω)u\in H^{\Bbbk_{u}}(\Omega). Therefore, the inf-sup condition eq. 38 holds. ∎

2.1.2. Convergence analysis

We proceed to the error analysis for the discrete nonlinear problem eq. 13. Let ρ(Ihu){vhWh:|Ihuvh|hρ}.\mathcal{B}_{\rho}(I_{h}u)\coloneqq\{v_{h}\in W_{h}:{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq\rho\}. We define the nonlinear map μh:WhWh\mu_{h}:W_{h}\to W_{h} by

𝒟𝒩hs(Ihu)μh(vh),wh=3Bs(Ihu,Ihu,vh,wh)+Ls(wh)Bs(vh,vh,vh,wh)\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\mu_{h}(v_{h}),w_{h}\rangle=3B^{s}(I_{h}u,I_{h}u,v_{h},w_{h})+L^{s}(w_{h})-B^{s}(v_{h},v_{h},v_{h},w_{h}) (39)

for vh,whWh,0v_{h},w_{h}\in W_{h,0}. Due to the weak coercivity property in theorem 2.8, the nonlinear map μh\mu_{h} is well-defined.

The existence and local uniqueness of the solution uhu_{h} to the discrete nonlinear problem eq. 13 will be proven via an application of Brouwer’s fixed point theorem, which necessitates the use of two auxiliary lemmas illustrating that (i) μh\mu_{h} maps from a ball to itself; and (ii) the map μh\mu_{h} is contracting.

Lemma 2.9.

(Mapping from a ball to itself) Let uu be a regular isolated solution of the continuous nonlinear weak problem eq. 11. For a sufficiently large ϵ\epsilon and a sufficiently small mesh size hh, there exists a positive constant R(h)>0R(h)>0 such that:

|vhIhu|hR(h)|μh(vh)Ihu|hR(h)vhWh,0.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}-I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq R(h)\Rightarrow{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mu_{h}(v_{h})-I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq R(h)\quad\forall v_{h}\in W_{h,0}.
Proof.

Note that the solution uH2(Ω)H01(Ω)u\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega) of eq. 11 satisfies the discrete weak formulation eq. 13 due to the consistency result theorem 2.3, that is to say, there holds that

Ahs(u,wh)+Phs(u,wh)+Bs(u,u,u,wh)+Cs(u,wh)=Ls(wh)whWh,0.A_{h}^{s}(u,w_{h})+P_{h}^{s}(u,w_{h})+B^{s}(u,u,u,w_{h})+C^{s}(u,w_{h})=L^{s}(w_{h})\quad\forall w_{h}\in W_{h,0}. (40)

By the linearity of 𝒟𝒩hs(Ihu),H2\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\cdot,\cdot\rangle_{H^{2}}, the definition eq. 39 of the nonlinear map μh\mu_{h} and eq. 40, we calculate

𝒟\displaystyle\langle\mathcal{D} 𝒩hs(Ihu)(Ihuμh(vh)),wh=𝒟𝒩hs(Ihu)Ihu,wh𝒟𝒩hs(Ihu)μh(vh),wh\displaystyle\mathcal{N}^{s}_{h}(I_{h}u)(I_{h}u-\mu_{h}(v_{h})),w_{h}\rangle=\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)I_{h}u,w_{h}\rangle-\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)\mu_{h}(v_{h}),w_{h}\rangle
=Ahs(Ihu,wh)+Phs(Ihu,wh)+3Bs(Ihu,Ihu,Ihu,wh)+Cs(Ihu,wh)\displaystyle={A_{h}^{s}(I_{h}u,w_{h})+P_{h}^{s}(I_{h}u,w_{h})}+3B^{s}(I_{h}u,I_{h}u,I_{h}u,w_{h})+C^{s}(I_{h}u,w_{h})
3Bs(Ihu,Ihu,vh,wh)+Bs(vh,vh,vh,wh)Ls(wh)\displaystyle\quad-3B^{s}(I_{h}u,I_{h}u,v_{h},w_{h})+B^{s}(v_{h},v_{h},v_{h},w_{h})-L^{s}(w_{h})
=Ahs(Ihuu,wh)+Phs(Ihuu,wh)𝔑1+Cs(Ihuu,wh)𝔑2+(Bs(Ihu,Ihu,Ihu,wh)Bs(u,u,u,wh))𝔑3\displaystyle=\underbrace{A^{s}_{h}(I_{h}u-u,w_{h})+P^{s}_{h}(I_{h}u-u,w_{h})}_{\eqqcolon\mathfrak{N}_{1}}+\underbrace{C^{s}(I_{h}u-u,w_{h})}_{\eqqcolon\mathfrak{N}_{2}}+\underbrace{\left(B^{s}(I_{h}u,I_{h}u,I_{h}u,w_{h})-B^{s}(u,u,u,w_{h})\right)}_{\eqqcolon\mathfrak{N}_{3}}
+(2Bs(Ihu,Ihu,Ihu,wh)3Bs(Ihu,Ihu,vh,wh)+Bs(vh,vh,vh,wh))𝔑4\displaystyle\quad+\underbrace{\left(2B^{s}(I_{h}u,I_{h}u,I_{h}u,w_{h})-3B^{s}(I_{h}u,I_{h}u,v_{h},w_{h})+B^{s}(v_{h},v_{h},v_{h},w_{h})\right)}_{\eqqcolon\mathfrak{N}_{4}}

In what follows, we give upper bounds for each 𝔑i,i=1,2,3,4\mathfrak{N}_{i},i=1,2,3,4. Using the boundedness of Ahs+Phs,CsA^{s}_{h}+P^{s}_{h},C^{s} and the interpolation estimate [9, Eq. (5.3)] in the ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}-norm, we obtain

𝔑1|Ihuu|h|wh|hhmin{deg1,𝕜u2}|wh|h,\displaystyle\mathfrak{N}_{1}\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h},
𝔑2|Ihuu|h|wh|hhmin{deg1,𝕜u2}|wh|h.\displaystyle\mathfrak{N}_{2}\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

We rearrange terms in 𝔑3\mathfrak{N}_{3} and use the boundedness result lemma 2.4 and the interpolation result [9, Eq. (5.3)] to obtain

𝔑3\displaystyle\mathfrak{N}_{3} =Bs(Ihuu,Ihuu,Ihu,wh)+2Bs(Ihuu,Ihuu,u,wh)+3Bs(u,u,Ihuu,wh)\displaystyle=B^{s}(I_{h}u-u,I_{h}u-u,I_{h}u,w_{h})+2B^{s}(I_{h}u-u,I_{h}u-u,u,w_{h})+3B^{s}(u,u,I_{h}u-u,w_{h})
(|Ihuu|h2|Ihu|h+|Ihuu|h2|u|h+u22Ihuu0)|wh|h\displaystyle\lesssim\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+\|u\|_{2}^{2}\|I_{h}u-u\|_{0}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}
(h2min{deg1,𝕜u2}+hmin{deg+1,𝕜u})|wh|h.\displaystyle\lesssim\left(h^{2\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+h^{\min\{\mathrm{deg}+1,\Bbbk_{u}\}}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

Let eI=vhIhue_{I}=v_{h}-I_{h}u. We use the definition of Bs(,,,)B^{s}(\cdot,\cdot,\cdot,\cdot) and its boundedness to deduce that

𝔑4\displaystyle\mathfrak{N}_{4} =a3Ω{2(Ihu)3wh3(Ihu)2vhwh+vh3wh}\displaystyle=a_{3}\int_{\Omega}\left\{2(I_{h}u)^{3}w_{h}-3(I_{h}u)^{2}v_{h}w_{h}+v_{h}^{3}w_{h}\right\}
=a3Ω{(vh2(Ihu)2)vhwh+2(Ihu)2(Ihuvh)wh}\displaystyle=a_{3}\int_{\Omega}\left\{\left(v_{h}^{2}-(I_{h}u)^{2}\right)v_{h}w_{h}+2(I_{h}u)^{2}(I_{h}u-v_{h})w_{h}\right\}
=a3Ω{eI(eI+2Ihu)(eI+Ihu)wh2(Ihu)2eIwh}\displaystyle=a_{3}\int_{\Omega}\left\{e_{I}(e_{I}+2I_{h}u)(e_{I}+I_{h}u)w_{h}-2(I_{h}u)^{2}e_{I}w_{h}\right\}
=a3Ω{eI(eI2+3eIIhu+2(Ihu)2)wh2(Ihu)2eIwh}\displaystyle=a_{3}\int_{\Omega}\left\{e_{I}\left(e_{I}^{2}+3e_{I}I_{h}u+2(I_{h}u)^{2}\right)w_{h}-2(I_{h}u)^{2}e_{I}w_{h}\right\}
=a3Ω(eI3+3eI3Ihu)wh=Bs(eI,eI,eI,wh)+3Bs(eI,eI,Ihu,wh)\displaystyle=a_{3}\int_{\Omega}\left(e_{I}^{3}+3e_{I}^{3}I_{h}u\right)w_{h}=B^{s}(e_{I},e_{I},e_{I},w_{h})+3B^{s}(e_{I},e_{I},I_{h}u,w_{h})
|eI|h2(|eI|h+|Ihu|h)|wh|h.\displaystyle\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{I}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{I}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

Hence, we combine the above bounds for 𝔑i\mathfrak{N}_{i}, i=1,2,3,4i=1,2,3,4 to have

D𝒩hs(Ihu)(Ihuμh(vh)),wh(hmin{deg1,𝕜u2}+hmin{2deg2,2𝕜u4,deg+1,𝕜u}+|eI|h2(|eI|h+1))|wh|h.\langle D\mathcal{N}^{s}_{h}(I_{h}u)(I_{h}u-\mu_{h}(v_{h})),w_{h}\rangle\lesssim\left(h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+h^{\min\{2\mathrm{deg}-2,2\Bbbk_{u}-4,\mathrm{deg}+1,\Bbbk_{u}\}}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{I}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{I}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+1\right)\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

By the inf-sup condition eq. 38 for the perturbed bilinear form, we further deduce that there exists a whWhw_{h}\in W_{h} with |wh|h=1{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}=1 such that |Ihuμh(vh)|hD𝒩hs(Ihu)(Ihuμh(vh)),wh.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-\mu_{h}(v_{h})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\langle D\mathcal{N}^{s}_{h}(I_{h}u)(I_{h}u-\mu_{h}(v_{h})),w_{h}\rangle. Since |eI|hR(h){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{I}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq R(h), we obtain

|Ihuμh(vh)|h(hmin{deg1,𝕜u2}+hmin{2deg2,2𝕜u4,deg+1,𝕜u}+R(h)2(R(h)+1))\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-\mu_{h}(v_{h})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\left(h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+h^{\min\{2\mathrm{deg}-2,2\Bbbk_{u}-4,\mathrm{deg}+1,\Bbbk_{u}\}}+R(h)^{2}\left(R(h)+1\right)\right)
{Cu(2hmin{deg1,𝕜u2}+R(h)2(1+R(h)))for 2deg3,𝕜u4,Cu(hmin{deg1,𝕜u2}+hmin{deg+1,2𝕜u4}+R(h)2(1+R(h)))for deg>3,𝕜u4.\displaystyle\quad\leq\begin{cases}C_{u}\left(2h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+R(h)^{2}(1+R(h))\right)&\text{for }2\leq\mathrm{deg}\leq 3,\Bbbk_{u}\leq 4,\\ C_{u}\left(h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+h^{\min\{\mathrm{deg}+1,2\Bbbk_{u}-4\}}+R(h)^{2}(1+R(h))\right)&\text{for }\mathrm{deg}>3,\Bbbk_{u}\leq 4.\end{cases}

Note that there are other cases when 𝕜u>4\Bbbk_{u}>4 and we only focus on the case of 𝕜u4\Bbbk_{u}\leq 4 here for brevity. The idea of the remainder of the proof is to choose an appropriate R(h)R(h) so that |Ihuμh(vh)|hR(h){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-\mu_{h}(v_{h})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq R(h). For simplicity of the calculation, we illustrate the case when 2deg3,𝕜u42\leq\mathrm{deg}\leq 3,\Bbbk_{u}\leq 4. To this end, we take R(h)=4Cuhmin{deg1,𝕜u2}R(h)=4C_{u}h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}} and choose hh satisfying h2min{deg1,𝕜u2}132Cu116.h^{2\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}\leq\frac{1}{32C_{u}}-\frac{1}{16}. This yields

|Ihuμh(vh)|h\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-\mu_{h}(v_{h})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} 2Cuhmin{deg1,𝕜u2}(1+CuR(h)2+Cu)\displaystyle\leq 2C_{u}h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}\left(1+C_{u}R(h)^{2}+C_{u}\right)
=2Cuhmin{deg1,𝕜u2}(1+32Cu3h2min{deg1,𝕜u2}+2Cu)R(h).\displaystyle=2C_{u}h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}\left(1+32C_{u}^{3}h^{2\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}+2C_{u}\right)\leq R(h).

This completes the proof. ∎

Lemma 2.10.

(Contraction result) For a sufficiently large ϵ\epsilon, a sufficiently small mesh size hh and any v1,v2R(h)(Ihu)v_{1},v_{2}\in\mathcal{B}_{R(h)}(I_{h}u), there holds

|μh(v1)μh(v2)|hhmin{deg1,𝕜u2}|v1v2|h.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mu_{h}(v_{1})-\mu_{h}(v_{2})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\mathrm{deg}-1,\Bbbk_{u}-2\}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{1}-v_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. (41)
Proof.

For whWhw_{h}\in W_{h}, we use the definition eq. 39 of the nonlinear map μh\mu_{h}, the definition eq. 37 and linearity of 𝒟𝒩hs(Ihu),\langle\mathcal{D}\mathcal{N}_{h}^{s}(I_{h}u)\cdot,\cdot\rangle to calculate

\displaystyle\langle 𝒟𝒩hs(Ihu)(μh(v1)μh(v2)),wh\displaystyle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)(\mu_{h}(v_{1})-\mu_{h}(v_{2})),w_{h}\rangle
=3Bs(Ihu,Ihu,v1,wh)Bs(v1,v1,v1,wh)3Bs(Ihu,Ihu,v2,wh)+Bs(v2,v2,v2,wh)\displaystyle=3B^{s}(I_{h}u,I_{h}u,v_{1},w_{h})-B^{s}(v_{1},v_{1},v_{1},w_{h})-3B^{s}(I_{h}u,I_{h}u,v_{2},w_{h})+B^{s}(v_{2},v_{2},v_{2},w_{h})
=a3Ω(3(Ihu)2v1whv13wh)a3Ω(3(Ihu)2v2whv23wh)\displaystyle=a_{3}\int_{\Omega}\left(3(I_{h}u)^{2}v_{1}w_{h}-v_{1}^{3}w_{h}\right)-a_{3}\int_{\Omega}\left(3(I_{h}u)^{2}v_{2}w_{h}-v_{2}^{3}w_{h}\right)
=a3Ω(((Ihu)2v12)v1wh+2(Ihu)2(v1v2)wh((Ihu)2v22)v2wh)\displaystyle=a_{3}\int_{\Omega}\left(\left((I_{h}u)^{2}-v_{1}^{2}\right)v_{1}w_{h}+2(I_{h}u)^{2}(v_{1}-v_{2})w_{h}-\left((I_{h}u)^{2}-v_{2}^{2}\right)v_{2}w_{h}\right)
=a3Ω((Ihuv1)(v1Ihu)(v1v2)wh+2(Ihuv1)Ihu(v1v2)wh+(Ihuv1)(Ihu+v1)v2wh)\displaystyle=a_{3}\int_{\Omega}((I_{h}u-v_{1})(v_{1}-I_{h}u)(v_{1}-v_{2})w_{h}+2(I_{h}u-v_{1})I_{h}u(v_{1}-v_{2})w_{h}+(I_{h}u-v_{1})(I_{h}u+v_{1})v_{2}w_{h})
+2a3Ω(Ihu(v1v2)(Ihuv2)wh+Ihu(v1v2)v2wh)a3Ω(Ihuv2)(Ihu+v2)v2wh\displaystyle\quad+2a_{3}\int_{\Omega}\left(I_{h}u(v_{1}-v_{2})(I_{h}u-v_{2})w_{h}+I_{h}u(v_{1}-v_{2})v_{2}w_{h}\right)-a_{3}\int_{\Omega}(I_{h}u-v_{2})(I_{h}u+v_{2})v_{2}w_{h}
=a3Ω(Ihuv1)(v1Ihu)(v1v2)wh+2a3Ω(Ihuv1)Ihu(v1v2)wh\displaystyle=a_{3}\int_{\Omega}(I_{h}u-v_{1})(v_{1}-I_{h}u)(v_{1}-v_{2})w_{h}+2a_{3}\int_{\Omega}(I_{h}u-v_{1})I_{h}u(v_{1}-v_{2})w_{h}
+2a3Ω(Ihuv2)Ihu(v1v2)wh+a3Ω(v1v2)((Ihuv1)+(Ihuv2))((v2Ihu)+Ihu)wh.\displaystyle\quad+2a_{3}\int_{\Omega}(I_{h}u-v_{2})I_{h}u(v_{1}-v_{2})w_{h}+a_{3}\int_{\Omega}(v_{1}-v_{2})\left((I_{h}u-v_{1})+(I_{h}u-v_{2})\right)\left((v_{2}-I_{h}u)+I_{h}u\right)w_{h}.

Let e1=Ihuv1e_{1}=I_{h}u-v_{1}, e2=Ihuv2e_{2}=I_{h}u-v_{2} and e=v1v2e=v_{1}-v_{2}. We make some elementary manipulations and use the boundedness of BsB^{s} and the inequality of geometric and arithmetic means to yield

\displaystyle\langle 𝒟𝒩hs(Ihu)(μh(v1)μh(v2)),wh\displaystyle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)(\mu_{h}(v_{1})-\mu_{h}(v_{2})),w_{h}\rangle
=a3Ω(e12)ewh+2a3Ωe1(Ihu)ewh+2a3Ωe2(Ihu)ewh+a3Ω{ewh(e1Ihu+e2Ihue1e2e22)}\displaystyle=a_{3}\int_{\Omega}(-e_{1}^{2})ew_{h}+2a_{3}\int_{\Omega}e_{1}(I_{h}u)ew_{h}+2a_{3}\int_{\Omega}e_{2}(I_{h}u)ew_{h}+a_{3}\int_{\Omega}\{ew_{h}(e_{1}I_{h}u+e_{2}I_{h}u-e_{1}e_{2}-e_{2}^{2})\}
(|e1|h2+|Ihu|h|e1|h+|e2|h|Ihu|h+|e1|h|e2|h+|e2|h2)|e|h|wh|h\displaystyle\lesssim\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}
(|e1|h2+|e2|h2+|e1|h+|e2|h)|e|h|wh|h(R(h)2+R(h))|e|h|wh|h.\displaystyle\lesssim\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{2}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\left(R(h)^{2}+R(h)\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}.

By the inf-sup condition eq. 38, we know that there exists whWhw_{h}\in W_{h} with |wh|h=1{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|w_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}=1 such that

βc2|μh(v1)μh(v2)|h𝒟𝒩hs(Ihu)(μh(v1)μh(v2)),wh.\frac{\beta_{c}}{2}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mu_{h}(v_{1})-\mu_{h}(v_{2})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\langle\mathcal{D}\mathcal{N}^{s}_{h}(I_{h}u)(\mu_{h}(v_{1})-\mu_{h}(v_{2})),w_{h}\rangle.

Therefore, we have |μh(v1)μh(v2)|hR(h)(1+R(h))|e|h.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mu_{h}(v_{1})-\mu_{h}(v_{2})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim R(h)(1+R(h)){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. Noting that R(h)(1+R(h))<1R(h)(1+R(h))<1 for 0<R(h)<120<R(h)<\frac{1}{2} completes the proof. ∎

The existence and local uniqueness of the discrete solution uhu_{h} can now be obtained via the application of Brouwer’s fixed point theorem [20].

Theorem 2.11.

(Convergence in ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}-norm) Let uu be a regular isolated solution of the nonlinear problem eq. 11. For a sufficiently large ϵ\epsilon and a sufficiently small hh, there exists a unique solution uhu_{h} of the discrete nonlinear problem eq. 13 within the local ball R(h)(Ihu)\mathcal{B}_{R(h)}(I_{h}u). Furthermore, we have |uuh|hhmin{deg1,𝕜u2}.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\deg-1,\Bbbk_{u}-2\}}.

Proof.

A use of lemma 2.9 yields that the nonlinear map μh\mu_{h} maps a closed convex set R(h)(Ihu)Wh\mathcal{B}_{R(h)}(I_{h}u)\subset W_{h} to itself. Moreover it is a contracting map. Therefore, an application of Brouwer’s fixed point theorem yields that μh\mu_{h} has at least one fixed point, say uhu_{h}, in this ball R(h)(Ihu)\mathcal{B}_{R(h)}(I_{h}u). The uniqueness of the solution to eq. 13 in that ball R(h)(Ihu)\mathcal{B}_{R(h)}(I_{h}u) follows from the contraction result in lemma 2.10. Meanwhile, we have by lemma 2.9 that

|uhIhu|hhmin{deg1,𝕜u2}.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u_{h}-I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\deg-1,\Bbbk_{u}-2\}}. (42)

The error estimate is then obtained straightforwardly using the triangle inequality |uuh|h|uIhu|h+|Ihuuh|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u-I_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} combined with eq. 42 and the interpolation estimate [9, Eq. (5.3)]. ∎

It follows from theorem 2.11 that optimal convergence rates are achieved in the mesh-dependent norm ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. This will be numerically verified in section 3.

2.1.3. Estimates in the L2L^{2}-norm

We derive an L2L^{2} error estimate using a duality argument in this subsection. To this end, we consider the following linear dual problem to the primal nonlinear problem eq. 9:

{2B((𝒟2χ))+a1χ+3a3u2χ=fdualin Ω,χ=0,𝒟2χ=𝟎on Ω,\begin{cases}2B\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}\chi))+a_{1}\chi+3a_{3}u^{2}\chi=f_{dual}\quad&\text{in }\Omega,\\ \chi=0,\quad\mathcal{D}^{2}\chi=\mathbf{0}\quad&\text{on }\partial\Omega,\end{cases} (43)

for fdualL2(Ω)f_{dual}\in L^{2}(\Omega). For smooth domains Ω\Omega, it can be deduced by a classical elliptic regularity result that χH4(Ω)\chi\in H^{4}(\Omega). The corresponding weak form is: find χH2(Ω)H01(Ω)\chi\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega) such that

2BΩ𝒟2χ:𝒟2v+a1Ωχv+3a3Ωu2χv=ΩfdualvvH2(Ω)H01(Ω),2B\int_{\Omega}\mathcal{D}^{2}\chi\colon\mathcal{D}^{2}v+a_{1}\int_{\Omega}\chi v+3a_{3}\int_{\Omega}u^{2}\chi v=\int_{\Omega}f_{dual}v\quad\forall v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega),

that is to say,

𝒟𝒩s(u)χ,vH2=𝒟𝒩hs(u)χ,v=(fdual,v)0.\langle\mathcal{D}\mathcal{N}^{s}(u)\chi,v\rangle_{H^{2}}=\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\chi,v\rangle=(f_{dual},v)_{0}. (44)
Remark 2.12.

The first equality in eq. 44 holds since uH2(Ω),χH2(Ω)u\in H^{2}(\Omega),\chi\in H^{2}(\Omega) and vH2(Ω)v\in H^{2}(\Omega).

We give two auxiliary results in the following.

Lemma 2.13.

For uH𝕜u(Ω)u\in H^{\Bbbk_{u}}(\Omega), 𝕜u>2\Bbbk_{u}>2, χH4(Ω)H01(Ω)\chi\in H^{4}(\Omega)\cap H^{1}_{0}(\Omega) and IhuWh,0H01(Ω)I_{h}u\in W_{h,0}\subset H^{1}_{0}(\Omega), there holds that

Ahs(Ihuu,χ)+Phs(Ihuu,χ)hmin{deg+1,𝕜u}χ4.A^{s}_{h}(I_{h}u-u,\chi)+P^{s}_{h}(I_{h}u-u,\chi)\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4}.
Proof.

Note that χ=0\llbracket\nabla\chi\rrbracket=0 since χH4(Ω)\chi\in H^{4}(\Omega) and χ=0\chi=0 on Ω\partial\Omega. We calculate

As(Ihuu,χ)+Phs(Ihuu,χ)\displaystyle A^{s}(I_{h}u-u,\chi)+P^{s}_{h}(I_{h}u-u,\chi) =ThT2B𝒟2(Ihuu):𝒟2χ2BeI{{2(Ihuu)ν2}}χ\displaystyle=\sum_{T\in\mathcal{E}_{h}}\int_{T}2B\mathcal{D}^{2}(I_{h}u-u)\colon\mathcal{D}^{2}\chi-2B\sum_{e\in\mathcal{E}_{I}}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}(I_{h}u-u)}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla\chi\rrbracket
2BeI{{2χν2}}(Ihuu)+eI2Bϵhe3e(Ihuu)χ\displaystyle\quad-2B\sum_{e\in\mathcal{E}_{I}}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}\chi}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla(I_{h}u-u)\rrbracket+\sum_{e\in\mathcal{E}_{I}}\frac{2B\epsilon}{h_{e}^{3}}\int_{e}\llbracket\nabla(I_{h}u-u)\rrbracket\llbracket\nabla\chi\rrbracket
=ThT2B𝒟2(Ihuu):𝒟2χ2BeI{{2χν2}}(Ihuu)\displaystyle=\sum_{T\in\mathcal{E}_{h}}\int_{T}2B\mathcal{D}^{2}(I_{h}u-u)\colon\mathcal{D}^{2}\chi-2B\sum_{e\in\mathcal{E}_{I}}\left\{\mskip-5.0mu\left\{\frac{\partial^{2}\chi}{\partial\nu^{2}}\right\}\mskip-5.0mu\right\}\llbracket\nabla(I_{h}u-u)\rrbracket
=ThT2B(Ihuu)((𝒟2χ))\displaystyle=\sum_{T\in\mathcal{E}_{h}}\int_{T}2B(I_{h}u-u)\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}\chi))
Ihuu0((𝒟2χ))0hmin{deg+1,𝕜u}χ4.\displaystyle\lesssim\|I_{h}u-u\|_{0}\|\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}\chi))\|_{0}\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4}.

Here, the last, second last, and third last steps follow from standard interpolation estimates, the Cauchy–Schwarz inequality, and integration by parts twice, respectively. ∎

Lemma 2.14.

The solution χ\chi of the linear dual problem eq. 43 belongs to H4(Ω)H^{4}(\Omega) on a smooth domain Ω\Omega and it holds that

χ4fdual0.\|\chi\|_{4}\lesssim\|f_{dual}\|_{0}. (45)
Proof.

We can use the inf-sup condition eq. 19 for the linear operator 𝒟𝒩s(u),\langle\mathcal{D}\mathcal{N}^{s}(u)\cdot,\cdot\rangle, the weak form eq. 44 and the Cauchy–Schwarz inequality to obtain

|χ|2sup|w|2=1wH2H01𝒟𝒩s(u)χ,wH2=sup|w|2=1wH2H01(fdual,w)0fdual0w0|w|2=1.|\chi|_{2}\lesssim\sup_{\overset{w\in H^{2}\cap H^{1}_{0}}{|w|_{2}=1}}\langle\mathcal{D}\mathcal{N}^{s}(u)\chi,w\rangle_{H^{2}}=\sup_{\overset{w\in H^{2}\cap H^{1}_{0}}{|w|_{2}=1}}(f_{dual},w)_{0}\lesssim\|f_{dual}\|_{0}\underbrace{\|w\|_{0}}_{\lesssim|w|_{2}=1}. (46)

By the form of eq. 44, the boundedness of Bs(u,u,,)B^{s}(u,u,\cdot,\cdot) and Cs(,)C^{s}(\cdot,\cdot), and eq. 46, we have

((𝒟2χ))0\displaystyle\|\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}\chi))\|_{0} =3Bs(u,u,χ,)Cs(χ,)+(fdual,)00χ0+fdual0fdual0.\displaystyle=\|-3B^{s}(u,u,\chi,\cdot)-C^{s}(\chi,\cdot)+(f_{dual},\cdot)_{0}\|_{0}\lesssim{\|\chi\|_{0}}+\|f_{dual}\|_{0}\lesssim\|f_{dual}\|_{0}. (47)

Using a bootstrapping argument in elliptic regularity (see, e.g., [18, Section 6.3]), we can deduce that χH4(Ω)\chi\in H^{4}(\Omega) in a smooth domain Ω\Omega. The regularity estimate eq. 45 follows from eq. 47. ∎

We are ready to derive the L2L^{2} a priori error estimates.

Theorem 2.15.

(L2L^{2} error estimate) Under the same conditions as theorem 2.11 and assuming further that deg>1\deg>1 (since the problem is fourth-order), the discrete solution uhu_{h} approximates uu such that

uuh0{hmin{deg+1,𝕜u}for deg3,h2min{deg1,𝕜u2}for deg=2.\|u-u_{h}\|_{0}\lesssim\begin{cases}h^{\min\{\deg+1,\Bbbk_{u}\}}&\text{for }\deg\geq 3,\\ h^{2\min\{\deg-1,\Bbbk_{u}-2\}}&\text{for }\deg=2.\end{cases} (48)
Proof.

Taking fdual=IhuuhWhH1(Ω)H2(𝒯h)f_{dual}=I_{h}u-u_{h}\in W_{h}\subset H^{1}(\Omega)\cap H^{2}(\mathcal{T}_{h}) in eq. 43 and multiplying eq. 43 by a test function vh=Ihuuhv_{h}=I_{h}u-u_{h} and integrating by parts, we obtain 𝒟𝒩hs(u)χ,Ihuuh=Ihuuh02.\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\chi,I_{h}u-u_{h}\rangle=\|I_{h}u-u_{h}\|_{0}^{2}. It follows from the fact that uH𝕜u(Ω)u\in H^{\Bbbk_{u}}(\Omega), 𝕜u4\Bbbk_{u}\geq 4, and the definition eq. 11 of the nonlinear continuous weak form 𝒩s(u)\mathcal{N}^{s}(u)\cdot that

Ihuuh02\displaystyle\|I_{h}u-u_{h}\|_{0}^{2} =𝒟𝒩hs(u)χ,Ihuuh+𝒩hs(uh)(Ihχ)𝒩hs(u)(Ihχ)\displaystyle=\langle\mathcal{D}\mathcal{N}^{s}_{h}(u)\chi,I_{h}u-u_{h}\rangle+\mathcal{N}^{s}_{h}(u_{h})(I_{h}\chi)-\mathcal{N}^{s}_{h}(u)(I_{h}\chi)
=Ahs(χ,Ihuuh)+Phs(χ,Ihuuh)+Cs(χ,Ihuuh)+3Bs(u,u,χ,Ihuuh)\displaystyle=A^{s}_{h}(\chi,I_{h}u-u_{h})+P^{s}_{h}(\chi,I_{h}u-u_{h})+C^{s}(\chi,I_{h}u-u_{h})+3B^{s}(u,u,\chi,I_{h}u-u_{h})
+Ahs(uh,Ihχ)+Phs(uh,Ihχ)+Cs(uh,Ihχ)+Bs(uh,uh,uh,Ihχ)\displaystyle\quad+A^{s}_{h}(u_{h},I_{h}\chi)+P^{s}_{h}(u_{h},I_{h}\chi)+C^{s}(u_{h},I_{h}\chi)+B^{s}(u_{h},u_{h},u_{h},I_{h}\chi)
Ahs(u,Ihχ)Phs(u,Ihχ)Cs(u,Ihχ)Bs(u,u,u,Ihχ)\displaystyle\quad-A^{s}_{h}(u,I_{h}\chi)-P^{s}_{h}(u,I_{h}\chi)-C^{s}(u,I_{h}\chi)-B^{s}(u,u,u,I_{h}\chi)
=Ahs(Ihuu,χ)+Ahs(uuh,χIhχ)+Phs(Ihuu,χ)+Phs(uuh,χIhχ)𝔘1\displaystyle=\underbrace{A^{s}_{h}(I_{h}u-u,\chi)+A^{s}_{h}(u-u_{h},\chi-I_{h}\chi)+P^{s}_{h}(I_{h}u-u,\chi)+P^{s}_{h}(u-u_{h},\chi-I_{h}\chi)}_{\eqqcolon\mathfrak{U}_{1}}
+Cs(Ihuu,χ)+Cs(uuh,χIhχ)𝔘2\displaystyle\quad+\underbrace{C^{s}(I_{h}u-u,\chi)+C^{s}(u-u_{h},\chi-I_{h}\chi)}_{\eqqcolon\mathfrak{U}_{2}}
+3Bs(u,u,Ihuuh,χIhχ)+3Bs(u,u,Ihuu,Ihχ)𝔘3\displaystyle\quad+\underbrace{3B^{s}(u,u,I_{h}u-u_{h},\chi-I_{h}\chi)+3B^{s}(u,u,I_{h}u-u,I_{h}\chi)}_{\eqqcolon\mathfrak{U}_{3}}
+Bs(uh,uh,uh,Ihχ)3Bs(u,u,uh,Ihχ)+2Bs(u,u,u,Ihχ)𝔘4.\displaystyle\quad+\underbrace{B^{s}(u_{h},u_{h},u_{h},I_{h}\chi)-3B^{s}(u,u,u_{h},I_{h}\chi)+2B^{s}(u,u,u,I_{h}\chi)}_{\eqqcolon\mathfrak{U}_{4}}.

We bound each 𝔘i\mathfrak{U}_{i} separately using the boundedness of Ahs,Phs,BsA^{s}_{h},P^{s}_{h},B^{s} and CsC^{s}, theorem 2.11 and standard interpolation estimates. This leads to

𝔘1hmin{deg+1,𝕜u}χ4+|uuh|h|χIhχ|hhmin{deg+1,𝕜u}χ4,\displaystyle\mathfrak{U}_{1}\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4}+{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\chi-I_{h}\chi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4},
𝔘2Ihuu0χ0+|uuh|h|χIhχ|hhmin{deg+1,𝕜u}χ4,\displaystyle\mathfrak{U}_{2}\lesssim{\|I_{h}u-u\|_{0}}{\|\chi\|_{0}}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\chi-I_{h}\chi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4},
𝔘3u22|Ihuuh|h|χIhχ|h+u22Ihuu0Ihχ0hmin{deg+1,𝕜u}χ4.\displaystyle\mathfrak{U}_{3}\lesssim\|u\|_{2}^{2}{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\chi-I_{h}\chi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}+\|u\|_{2}^{2}{\|I_{h}u-u\|_{0}}{\|I_{h}\chi\|_{0}}\lesssim h^{\min\{\deg+1,\Bbbk_{u}\}}\|\chi\|_{4}.

Setting e3=uhue_{3}=u_{h}-u and estimating 𝔘4\mathfrak{U}_{4} as in 4\mathfrak{R}_{4} of lemma 2.9 with the use of theorem 2.11 and |Ihχ|hχ2χ4{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}\chi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\lesssim\|\chi\|_{2}\leq\|\chi\|_{4} yields

𝔘4|e3|h2(|e3|h+|u|h)|Ihχ|hh2min{deg1,𝕜u2}(hmin{deg1,𝕜u2}+1)χ4.\mathfrak{U}_{4}\lesssim{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{3}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}^{2}\left({\left|\kern-1.07639pt\left|\kern-1.07639pt\left|e_{3}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}\right){{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|I_{h}\chi\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}}\lesssim h^{2\min\{\deg-1,\Bbbk_{u}-2\}}(h^{\min\{\deg-1,\Bbbk_{u}-2\}}+1)\|\chi\|_{4}.

Combining the above estimates for 𝔘i\mathfrak{U}_{i} (i=1,2,3,4i=1,2,3,4) and using the regularity estimate eq. 45 and χ4Ihuuh0\|\chi\|_{4}\lesssim\|I_{h}u-u_{h}\|_{0}, we obtain

Ihuuh0{hmin{deg+1,𝕜u}if deg3,h2min{deg1,𝕜u2}if deg=2.\|I_{h}u-u_{h}\|_{0}\lesssim\begin{cases}h^{\min\{\deg+1,\Bbbk_{u}\}}&\text{if }\deg\geq 3,\\ h^{2\min\{\deg-1,\Bbbk_{u}-2\}}&\text{if }\deg=2.\end{cases}

Hence, eq. 48 follows from the triangle inequality and standard interpolation estimates. ∎

theorem 2.15 implies that for quadratic approximations to the sufficiently regular solution of eq. 9, there is a sub-optimal convergence rate in the L2L^{2}-norm while for higher order (3\geq 3) approximations, we expect optimal L2L^{2} error rates. We shall see numerical verifications of this in the subsequent sections.

2.1.4. The inconsistent discrete form

The above analysis considers the consistent weak formulation eq. 13. In practice, Xia et al. [31] adopted the inconsistent discrete weak form in the implementation due to its cheaper assembly cost: find uhWh,bu_{h}\in W_{h,b} such that

𝒩~hs(uh)th=A~hs(uh,th)+Bs(uh,uh,uh,th)+Cs(uh,th)+Phs(uh,th)=0thWh,0,\tilde{\mathcal{N}}_{h}^{s}(u_{h})t_{h}=\tilde{A}^{s}_{h}(u_{h},t_{h})+B^{s}(u_{h},u_{h},u_{h},t_{h})+C^{s}(u_{h},t_{h})+P^{s}_{h}(u_{h},t_{h})=0\quad\forall t_{h}\in W_{h,0}, (49)

where A~hs(u,t)2BT𝒯hT𝒟2u:𝒟2t.\tilde{A}^{s}_{h}(u,t)\coloneqq 2B\sum_{T\in\mathcal{T}_{h}}\int_{T}\mathcal{D}^{2}u\colon\mathcal{D}^{2}t. Comparing A~hs\tilde{A}^{s}_{h} and AhsA^{s}_{h}, the missing terms are the interior facet integrals arising from piecewise integration by parts and symmetrisation. Due to the absence of these terms in A~hs\tilde{A}^{s}_{h}, one can immediately notice that the discrete weak formulation eq. 49 is inconsistent in the sense that the solution uu of the strong form eq. 9 does not satisfy the weak form eq. 49, as opposed to the result of theorem 2.3.

Despite this inconsistency, in practice this also leads to a convergent numerical scheme with similar convergence rates, as illustrated in section 3. This is not surprising; a similar idea has also been applied and introduced as weakly over-penalised symmetric interior penalty (WOPSIP) methods in [10] for second-order elliptic PDEs and in [8] for biharmonic equations.

Remark 2.16.

The excessive size of the penalty parameter in the WOPSIP method could induce ill-conditioned linear systems. No such effects are observed in our numerical results.

2.2. A priori error estimates for (𝒫2)(\mathcal{P}2)

Problem (𝒫2)(\mathcal{P}2) is a special form of the classical LdG model of nematic LC. Finite element analysis for a more general form using conforming discretisations has been studied in [13, 14]. More specifically, Davis and Gartland [14] gave an abstract nonlinear finite element convergence analysis where an optimal H1H^{1} error bound is proved on convex domains with piecewise linear polynomial approximations, but do not derive an error bound in the L2L^{2} norm. Recently, Maity, Majumdar and Nataraj [21] analysed the discontinuous Galerkin finite element method for a two-dimensional reduced LdG free energy, where optimal a priori error estimates in the L2L^{2}-norm with exact solutions in H2H^{2} are achieved for a piecewise linear discretisation. Both works only focus on piecewise linear approximations. In this section, we will follow similar steps to section 2.1 to prove the H1H^{1}- and L2L^{2}-convergence rates for the problem (𝒫2)(\mathcal{P}2) with the use of common continuous Lagrange elements of arbitrary positive degree. Since the approach is similar to the previous subsections, we omit some details for brevity.

The continuous weak formulation of (𝒫2)(\mathcal{P}2) in two dimensions (the three-dimensional case can be tackled similarly) is given by: find 𝐐Hb1(Ω,S0)\mathbf{Q}\in H^{1}_{b}(\Omega,S_{0}) such that

𝒩n(𝐐)𝐏An(𝐐,𝐏)+Bn(𝐐,𝐐,𝐐,𝐏)+Cn(𝐐,𝐏)=0𝐏𝐇01(Ω),\mathcal{N}^{n}(\mathbf{Q})\mathbf{P}\coloneqq A^{n}(\mathbf{Q},\mathbf{P})+B^{n}(\mathbf{Q},\mathbf{Q},\mathbf{Q},\mathbf{P})+C^{n}(\mathbf{Q},\mathbf{P})=0\quad\forall\mathbf{P}\in\mathbf{H}^{1}_{0}(\Omega), (50)

where the bilinear forms are An(𝐐,𝐏)KΩ𝐐 . . . 𝐏,Cn(𝐐,𝐏)2lΩ𝐐:𝐏,A^{n}(\mathbf{Q},\mathbf{P})\coloneqq K\int_{\Omega}\nabla\mathbf{Q}\mathrel{\mathchoice{\vbox to4.30554pt{\hbox{$\displaystyle.$}\vss\hbox{$\displaystyle.$}\vss\hbox{$\displaystyle.$}}}{\vbox to4.30554pt{\hbox{$\textstyle.$}\vss\hbox{$\textstyle.$}\vss\hbox{$\textstyle.$}}}{\vbox to3.01389pt{\hbox{$\scriptstyle.$}\vss\hbox{$\scriptstyle.$}\vss\hbox{$\scriptstyle.$}}}{\vbox to2.15277pt{\hbox{$\scriptscriptstyle.$}\vss\hbox{$\scriptscriptstyle.$}\vss\hbox{$\scriptscriptstyle.$}}}}\nabla\mathbf{P},\ C^{n}(\mathbf{Q},\mathbf{P})\coloneqq-2l\int_{\Omega}\mathbf{Q}:\mathbf{P}, and the nonlinear operator is given by

Bn(Ψ,Φ,Θ,Ξ)4l3Ω((Ψ:Φ)(Θ:Ξ)+2(Ψ:Θ)(Φ:Ξ)).B^{n}(\Psi,\Phi,\Theta,\Xi)\coloneqq\frac{4l}{3}\int_{\Omega}\left((\Psi:\Phi)(\Theta:\Xi)+2(\Psi:\Theta)(\Phi:\Xi)\right). (51)

Since eq. 50 is nonlinear, we need to approximate the solution of its linearised version, i.e., find Θ𝐇01(Ω)\Theta\in\mathbf{H}^{1}_{0}(\Omega) such that

𝒟𝒩n(𝐐)Θ,ΦAn(Θ,Φ)+3Bn(𝐐,𝐐,Θ,Φ)+Cn(Θ,Φ)=𝒩n(𝐐)ΦΦ𝐇01(Ω),\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\Theta,\Phi\rangle\coloneqq A^{n}(\Theta,\Phi)+3B^{n}(\mathbf{Q},\mathbf{Q},\Theta,\Phi)+C^{n}(\Theta,\Phi)=-\mathcal{N}^{n}(\mathbf{Q})\Phi\quad\forall\Phi\in\mathbf{H}^{1}_{0}(\Omega), (52)

where ,\langle\cdot,\cdot\rangle represents the dual pairing between 𝐇1(Ω)\mathbf{H}^{-1}(\Omega) and 𝐇01(Ω)\mathbf{H}^{1}_{0}(\Omega).

Suppose 𝐐h𝐕h\mathbf{Q}_{h}\in\mathbf{V}_{h} approximates the solution of eq. 50 with the conforming finite element method on a finite dimensional space 𝐕h{𝐏𝐇1(Ω):𝐏deg(T),deg1,T𝒯h}\mathbf{V}_{h}\coloneqq\{\mathbf{P}\in\mathbf{H}^{1}(\Omega):\mathbf{P}\in\mathbb{Q}_{\deg}(T),\deg\geq 1,\forall T\in\mathcal{T}_{h}\}. Throughout this subsection we take deg1\deg\geq 1. Furthermore, we denote 𝐕h,0{𝐏𝐕h:𝐏=𝟎 on Ω}\mathbf{V}_{h,0}\coloneqq\{\mathbf{P}\in\mathbf{V}_{h}:\mathbf{P}=\mathbf{0}\text{ on }\partial\Omega\} and 𝐕h,b{𝐏𝐕h:𝐏=𝐐b on Ω}\mathbf{V}_{h,b}\coloneqq\{\mathbf{P}\in\mathbf{V}_{h}:\mathbf{P}=\mathbf{Q}_{b}\text{ on }\partial\Omega\}. We assume that the minimiser 𝐐\mathbf{Q} to be approximated is isolated, i.e., the linearised operator 𝒟𝒩n(𝐐),\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\cdot,\cdot\rangle is nonsingular. This is equivalent to the following continuous inf-sup condition [21, Eq. (2.8)]:

0<βQinfΘ𝐇01(Ω)Θ1=1supΦ𝐇01(Ω)Φ1=1𝒟𝒩n(𝐐)Θ,Φ=infΦ𝐇01(Ω)Φ1=1supΘ𝐇01(Ω)Θ1=1𝒟𝒩n(𝐐)Θ,Φ.0<\beta_{Q}\coloneqq\adjustlimits{\inf}_{\underset{\|\Theta\|_{1}=1}{\Theta\in\mathbf{H}^{1}_{0}(\Omega)}}{\sup}_{\underset{\|\Phi\|_{1}=1}{\Phi\in\mathbf{H}^{1}_{0}(\Omega)}}\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\Theta,\Phi\rangle=\adjustlimits{\inf}_{\underset{\|\Phi\|_{1}=1}{\Phi\in\mathbf{H}^{1}_{0}(\Omega)}}{\sup}_{\underset{\|\Theta\|_{1}=1}{\Theta\in\mathbf{H}^{1}_{0}(\Omega)}}\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\Theta,\Phi\rangle. (53)

With this inf-sup condition for 𝒟𝒩n(𝐐),\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\cdot,\cdot\rangle, we can obtain a stability result for the perturbed bilinear form 𝒟𝒩n(Ih𝐐),\langle\mathcal{D}\mathcal{N}^{n}(I_{h}\mathbf{Q})\cdot,\cdot\rangle by following similar steps as in the proof of theorem 2.8.

Theorem 2.17.

(Stability of the perturbed bilinear form) Let 𝐐\mathbf{Q} be a regular isolated solution of the nonlinear continuous weak form eq. 50 and Ih𝐐I_{h}\mathbf{Q} the interpolant of 𝐐\mathbf{Q}. For a sufficiently small mesh size hh, the following discrete inf-sup condition holds:

0<βQ2infΘ𝐇01(Ω)Θ1=1supΦ𝐇01(Ω)Φ1=1𝒟𝒩n(Ih𝐐)Θ,Φ.0<\frac{\beta_{Q}}{2}\leq\adjustlimits{\inf}_{\underset{\|\Theta\|_{1}=1}{\Theta\in\mathbf{H}^{1}_{0}(\Omega)}}{\sup}_{\underset{\|\Phi\|_{1}=1}{\Phi\in\mathbf{H}^{1}_{0}(\Omega)}}\langle\mathcal{D}\mathcal{N}^{n}(I_{h}\mathbf{Q})\Theta,\Phi\rangle. (54)

We give some auxiliary results about the operators An(,)A^{n}(\cdot,\cdot), Bn(,,,)B^{n}(\cdot,\cdot,\cdot,\cdot) and Cn(,)C^{n}(\cdot,\cdot) that can be verified via the Cauchy–Schwarz inequality, the Poincaré inequality, and Sobolev embeddings.

Lemma 2.18.

(Boundedness and coercivity of An(,)A^{n}(\cdot,\cdot)) For Θ,Φ𝐇01(Ω)\Theta,\Phi\in\mathbf{H}^{1}_{0}(\Omega), there holds

An(Θ,Φ)Θ1Φ1 and Θ12An(Θ,Θ).A^{n}(\Theta,\Phi)\lesssim\|\Theta\|_{1}\|\Phi\|_{1}\text{ and }\|\Theta\|_{1}^{2}\lesssim A^{n}(\Theta,\Theta).
Lemma 2.19.

(Boundedness of Bn(,,,)B^{n}(\cdot,\cdot,\cdot,\cdot), Cn(,)C^{n}(\cdot,\cdot)) For Ψ,Φ,Θ,Ξ𝐇1(Ω)\Psi,\Phi,\Theta,\Xi\in\mathbf{H}^{1}(\Omega), there holds

Bn(Ψ,Φ,Θ,Ξ)Ψ1Φ1Θ1Ξ1,Cn(Ψ,Φ)Ψ1Φ1,B^{n}(\Psi,\Phi,\Theta,\Xi)\lesssim\|\Psi\|_{1}\|\Phi\|_{1}\|\Theta\|_{1}\|\Xi\|_{1},\quad C^{n}(\Psi,\Phi)\lesssim\|\Psi\|_{1}\|\Phi\|_{1}, (55)

and for Ψ,Φ𝐇𝕜(Ω)\Psi,\Phi\in\mathbf{H}^{\Bbbk}(\Omega), 𝕜2\Bbbk\geq 2, Θ,Ξ𝐇1(Ω)\Theta,\Xi\in\mathbf{H}^{1}(\Omega),

Bn(Ψ,Φ,Θ,Ξ)Ψ𝕜Φ𝕜Θ1Ξ1.B^{n}(\Psi,\Phi,\Theta,\Xi)\lesssim\|\Psi\|_{\Bbbk}\|\Phi\|_{\Bbbk}\|\Theta\|_{1}\|\Xi\|_{1}. (56)

To proceed to error estimates for the nonlinear problem eq. 50, we define the nonlinear map ψ:𝐕h𝐕h\psi:\mathbf{V}_{h}\to\mathbf{V}_{h} by

𝒟𝒩n(Ih𝐐)ψ(Θh),Φh=3Bn(Ih𝐐,Ih𝐐,Θh,Φh)Bn(Θh,Θh,Θh,Φh)\langle\mathcal{D}\mathcal{N}^{n}(I_{h}\mathbf{Q})\psi(\Theta_{h}),\Phi_{h}\rangle=3B^{n}(I_{h}\mathbf{Q},I_{h}\mathbf{Q},\Theta_{h},\Phi_{h})-B^{n}(\Theta_{h},\Theta_{h},\Theta_{h},\Phi_{h})

for Θh,Φh𝐕h,0\Theta_{h},\Phi_{h}\in\mathbf{V}_{h,0}. Due to the stability result of theorem 2.17, the nonlinear map ψ\psi is well-defined. We define the local ball ρ(Ih𝐐){𝐏h𝐕h:Ih𝐐𝐏h1ρ}\mathcal{B}_{\rho}(I_{h}\mathbf{Q})\coloneqq\{\mathbf{P}_{h}\in\mathbf{V}_{h}:\|I_{h}\mathbf{Q}-\mathbf{P}_{h}\|_{1}\leq\rho\}. The following two auxiliary lemmas provide the necessary components for the application of Brouwer’s fixed point theorem.

Lemma 2.20.

(Mapping from a ball to itself) Let 𝐐\mathbf{Q} be a regular isolated solution of the continuous nonlinear weak problem eq. 50. For a sufficiently small mesh size hh, there exists a positive constant r(h)>0r(h)>0 such that:

𝐏hIh𝐐1r(h)ψ(𝐏h)Ih𝐐1r(h)𝐏h𝐕h,0.\|\mathbf{P}_{h}-I_{h}\mathbf{Q}\|_{1}\leq r(h)\Rightarrow\|\psi(\mathbf{P}_{h})-I_{h}\mathbf{Q}\|_{1}\leq r(h)\quad\forall\mathbf{P}_{h}\in\mathbf{V}_{h,0}.
Remark 2.21.

In fact, the choice of r(h)r(h) can be taken as r(h)=𝒪(hmin{deg,𝕜Q1})r(h)=\mathcal{O}(h^{\min\{\deg,\Bbbk_{Q}-1\}}) in the proof of lemma 2.20. Here, 𝕜Q2\Bbbk_{Q}\geq 2 denotes the regularity index of 𝐐\mathbf{Q}, i.e., 𝐐𝐇𝕜Q(Ω)\mathbf{Q}\in\mathbf{H}^{\Bbbk_{Q}}(\Omega).

Lemma 2.22.

(Contraction result) For a sufficiently small mesh size hh and any 𝐏1,𝐏2r(h)(Ih𝐐)\mathbf{P}_{1},\mathbf{P}_{2}\in\mathcal{B}_{r(h)}(I_{h}\mathbf{Q}), there holds

ψ(𝐏1)ψ(𝐏2)1hmin{deg,𝕜Q1}𝐏1𝐏21.\|\psi(\mathbf{P}_{1})-\psi(\mathbf{P}_{2})\|_{1}\lesssim h^{\min\{\deg,\Bbbk_{Q}-1\}}\|\mathbf{P}_{1}-\mathbf{P}_{2}\|_{1}. (57)
Remark 2.23.

In the proof of lemma 2.22, we have particularly used the stability property of the perturbed bilinear form as given by theorem 2.17.

Hence, the existence and local uniqueness of the discrete solution 𝐐h\mathbf{Q}_{h} can be derived by following similar steps as in the proof of theorem 2.11.

Theorem 2.24.

(Convergence in 1\|\cdot\|_{1}-norm) Let 𝐐\mathbf{Q} be a regular isolated solution of the nonlinear problem eq. 50. For a sufficiently small hh, there exists a unique solution 𝐐h\mathbf{Q}_{h} of the discrete nonlinear problem eq. 50 within the local ball r(h)(Ih𝐐)\mathcal{B}_{r(h)}(I_{h}\mathbf{Q}). Furthermore, we have 𝐐𝐐h1hmin{deg,𝕜Q1}.\|\mathbf{Q}-\mathbf{Q}_{h}\|_{1}\lesssim h^{\min\{\deg,\Bbbk_{Q}-1\}}.

We again employ an Aubin–Nitsche duality argument to derive L2L^{2} error estimates. To this end, we consider the following linear dual problem to the primal nonlinear problem eq. 10: find 𝐍𝐇01(Ω)\mathbf{N}\in\mathbf{H}^{1}_{0}(\Omega) such that

{KΔ𝐍+4l|𝐐|2𝐍+8l(𝐐:𝐍)𝐐2l𝐍=𝐆in Ω,𝐍=𝟎on Ω,\begin{cases}-K\Delta\mathbf{N}+4l|\mathbf{Q}|^{2}\mathbf{N}+8l(\mathbf{Q}:\mathbf{N})\mathbf{Q}-2l\mathbf{N}=\mathbf{G}&\text{in }\Omega,\\ \mathbf{N}=\mathbf{0}&\text{on }\partial\Omega,\end{cases} (58)

for a given 𝐆𝐋2(Ω)\mathbf{G}\in\mathbf{L}^{2}(\Omega) (we will make a particular choice for 𝐆\mathbf{G} in the proof of theorem 2.27). The weak form of eq. 58 is to find 𝐍𝐇01(Ω)\mathbf{N}\in\mathbf{H}^{1}_{0}(\Omega) such that

𝒟𝒩n(𝐐)𝐍,Φ=An(𝐍,Φ)+3Bn(𝐐,𝐐,𝐍,Φ)+Cn(𝐍,Φ)=(𝐆,Φ)0Φ𝐇01(Ω).\langle\mathcal{D}\mathcal{N}^{n}(\mathbf{Q})\mathbf{N},\Phi\rangle=A^{n}(\mathbf{N},\Phi)+3B^{n}(\mathbf{Q},\mathbf{Q},\mathbf{N},\Phi)+C^{n}(\mathbf{N},\Phi)=(\mathbf{G},\Phi)_{0}\quad\forall\Phi\in\mathbf{H}^{1}_{0}(\Omega). (59)

To derive the L2L^{2} a priori error estimates, we need two more auxiliary results.

Lemma 2.25.

For 𝐐𝐇𝕜Q(Ω)𝐇b1(Ω)\mathbf{Q}\in\mathbf{H}^{\Bbbk_{Q}}(\Omega)\cap\mathbf{H}^{1}_{b}(\Omega), 𝕜Q2\Bbbk_{Q}\geq 2, 𝐍𝐇2(Ω)𝐇01(Ω)\mathbf{N}\in\mathbf{H}^{2}(\Omega)\cap\mathbf{H}^{1}_{0}(\Omega) and Ih𝐐𝐕h𝐇b1(Ω)I_{h}\mathbf{Q}\in\mathbf{V}_{h}\subset\mathbf{H}^{1}_{b}(\Omega), it holds that

An(Ih𝐐𝐐,𝐍)hmin{deg+1,𝕜Q}𝐐𝕜Q𝐍2.A^{n}(I_{h}\mathbf{Q}-\mathbf{Q},\mathbf{N})\lesssim h^{\min\{\deg+1,\Bbbk_{Q}\}}\|\mathbf{Q}\|_{\Bbbk_{Q}}\|\mathbf{N}\|_{2}.
Lemma 2.26.

(Boundedness of the dual solution in the H2H^{2}-norm) The solution 𝐍\mathbf{N} to the weak form eq. 59 of the dual linear problem belongs to 𝐇2(Ω)𝐇01(Ω)\mathbf{H}^{2}(\Omega)\cap\mathbf{H}^{1}_{0}(\Omega) and it holds that

𝐍2𝐆0.\|\mathbf{N}\|_{2}\lesssim\|\mathbf{G}\|_{0}. (60)

Finally, we are ready to deduce an optimal L2L^{2} error estimate.

Theorem 2.27.

(L2L^{2} error estimate) Let 𝐐\mathbf{Q} be a regular solution of the nonlinear weak problem eq. 50 and 𝐐h\mathbf{Q}_{h} be the approximate solution to the discrete problem (having the same weak formulation as eq. 50). Then

𝐐𝐐h0hmin{deg+1,𝕜Q}(2+(3+h+h2+hmin{deg,𝕜Q1}+hmin{deg+1,𝕜Q})𝐐𝕜Q2)𝐐𝕜Q.\|\mathbf{Q}-\mathbf{Q}_{h}\|_{0}\lesssim h^{\min\{\deg+1,\Bbbk_{Q}\}}\left(2+\left(3+h+h^{2}+h^{\min\{\deg,\Bbbk_{Q}-1\}}+h^{\min\{\deg+1,\Bbbk_{Q}\}}\right)\|\mathbf{Q}\|_{\Bbbk_{Q}}^{2}\right)\|\mathbf{Q}\|_{\Bbbk_{Q}}. (61)

We will verify these results in the next section.

3. Numerical experiments

The proceeding section presents some a priori error estimates for both 𝐐\mathbf{Q} and uu in the decoupled case q=0q=0. We now test the convergence rate of the finite element approximations by the method of manufactured solutions (MMS) and experimentally investigate the coupled case q0q\neq 0 in two dimensions. To this end, we choose a nontrivial solution for each state variable and add an appropriate source term to the equilibrium equations, thus modifying the energy accordingly. We can then compute the numerical convergence order.

3.1. Test 1: on the unit square

In this test, the numerical runs are performed on the unit square Ω=(0,1)2\Omega=(0,1)^{2} and we take the following exact expressions for each state variable,

Q11e\displaystyle Q_{11}^{e} =(cos(π(2y1)(2x1)8))212,\displaystyle=\left(\cos\left(\frac{\pi(2y-1)(2x-1)}{8}\right)\right)^{2}-\frac{1}{2}, (62)
Q12e\displaystyle Q_{12}^{e} =cos(π(2y1)(2x1)8)sin(π(2y1)(2x1)8),\displaystyle=\cos\left(\frac{\pi(2y-1)(2x-1)}{8}\right)\sin\left(\frac{\pi(2y-1)(2x-1)}{8}\right),
ue\displaystyle u^{e} =10((x1)x(y1)y)3.\displaystyle=10\left((x-1)x(y-1)y\right)^{3}.

Then, in conducting the MMS, we are to solve the following governing equations

{4Bq4u2Q11+2Bq2u(x2uy2u)2KΔQ114lQ11+16lQ11(Q112+Q122)=𝔰1,4Bq4u2Q12+4Bq2u(xyu)2KΔQ124lQ12+16lQ12(Q112+Q122)=𝔰2,a1u+a2u2+a3u3+2B((𝒟2u))+Bq4(4(Q112+Q122)+1)u+2Bq2(t1+t2)=𝔰3,\begin{cases}4Bq^{4}u^{2}Q_{11}+2Bq^{2}u\left(\partial_{x}^{2}u-\partial_{y}^{2}u\right)-2K\Delta Q_{11}-4lQ_{11}+16lQ_{11}\left(Q_{11}^{2}+Q_{12}^{2}\right)=\mathfrak{s}_{1},\\ 4Bq^{4}u^{2}Q_{12}+4Bq^{2}u\left(\partial_{x}\partial_{y}u\right)-2K\Delta Q_{12}-4lQ_{12}+16lQ_{12}\left(Q_{11}^{2}+Q_{12}^{2}\right)=\mathfrak{s}_{2},\\ a_{1}u+a_{2}u^{2}+a_{3}u^{3}+2B\nabla\cdot(\nabla\cdot(\mathcal{D}^{2}u))+Bq^{4}\left(4\left(Q_{11}^{2}+Q_{12}^{2}\right)+1\right)u+2Bq^{2}(t_{1}+t_{2})=\mathfrak{s}_{3},\end{cases}

subject to Dirichlet boundary conditions for both uu and 𝐐\mathbf{Q} and a natural boundary condition for uu. Here, source terms 𝔰1\mathfrak{s}_{1}, 𝔰2\mathfrak{s}_{2} and 𝔰3\mathfrak{s}_{3} are derived by substituting eq. 62 to the left hand sides, and t1t_{1} and t2t_{2} are given by

t1(Q11+1/2)x2u+(Q11+1/2)y2u+2Q12xyu,t2x2(u(Q11+1/2))+y2(u(Q11+1/2))+2xy(uQ12).\begin{split}t_{1}&\coloneqq(Q_{11}+1/2)\partial_{x}^{2}u+(-Q_{11}+1/2)\partial_{y}^{2}u+2Q_{12}\partial_{x}\partial_{y}u,\\ t_{2}&\coloneqq\partial_{x}^{2}\left(u\left(Q_{11}+1/2\right)\right)+\partial_{y}^{2}(u(-Q_{11}+1/2))+2\partial_{x}\partial_{y}(uQ_{12}).\end{split}

We partition the domain Ω=(0,1)2\Omega=(0,1)^{2} into N×NN\times N squares with uniform mesh size h=1Nh=\frac{1}{N} (N=6N=6, 1212, 2424, 4848) and denote the numerical solutions by uhu_{h}, Q11,hQ_{11,h} and Q12,hQ_{12,h}. The numerical errors of uu and 𝐐\mathbf{Q} in the 0\|\cdot\|_{0}-, 1\|\cdot\|_{1}- and ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}-norms are defined as

eu0=ueuh0,eu1=ueuh1,|eu|h=|ueuh|h,\displaystyle\|\textbf{e}_{u}\|_{0}=\|u^{e}-u_{h}\|_{0},\quad\|\textbf{e}_{u}\|_{1}=\|u^{e}-u_{h}\|_{1},\quad{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u^{e}-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h},
e𝐐0=\displaystyle\|\textbf{e}_{\mathbf{Q}}\|_{0}= (Q11e,Q12e)(Q11,h,Q12,h)0,e𝐐1=(Q11e,Q12e)(Q11,h,Q12,h)1.\displaystyle\|(Q_{11}^{e},Q_{12}^{e})-(Q_{11,h},Q_{12,h})\|_{0},\quad\|\textbf{e}_{\mathbf{Q}}\|_{1}=\|(Q_{11}^{e},Q_{12}^{e})-(Q_{11,h},Q_{12,h})\|_{1}.

The convergence order is then calculated from the formula log2(errorh/2errorh).\log_{2}\left(\frac{\text{error}_{h/2}}{\text{error}_{h}}\right). Throughout this section, we use the parameter values a1=10a_{1}=-10, a2=0a_{2}=0, a3=10a_{3}=10, B=105B=10^{-5}, K=0.3K=0.3 and l=30l=30, similar to the simulations of oily streaks in [31].

Remark 3.1.

Since this is purely a numerical verification exercise, the manufactured solution can be physically unrealistic. However, we must specify a reasonable initial guess for Newton’s method, due to the nonlinearity of the problem. The initial guess throughout this section is taken to be (12(exact solution)+109)\left(\frac{1}{2}(\text{exact solution})+10^{-9}\right).

3.1.1. Convergence rate for q=0q=0

For 𝐐\mathbf{Q} we expect both optimal H1H^{1} and L2L^{2} rates, as illustrated in theorems 2.27 and 2.24. table 1 presents the numerical convergence rate for the finite elements [1]2[\mathbb{Q}_{1}]^{2}, [2]2[\mathbb{Q}_{2}]^{2} and [3]2[\mathbb{Q}_{3}]^{2}. Optimal L2L^{2} and H1H^{1} rates are shown with all choices of finite elements, as predicted.

N=1hN=\frac{1}{h} e𝐐0\|\textbf{e}_{\mathbf{Q}}\|_{0} rate e𝐐1\|\textbf{e}_{\mathbf{Q}}\|_{1} rate
[1]2[\mathbb{Q}_{1}]^{2} 6 8.12 ×104\times 10^{-4} 3.78 ×102\times 10^{-2}
12 2.02 ×104\times 10^{-4} 2.01 1.88 ×102\times 10^{-2} 1.01
24 5.05 ×105\times 10^{-5} 2.00 9.39 ×103\times 10^{-3} 1.00
48 1.26 ×105\times 10^{-5} 2.00 4.69 ×103\times 10^{-3} 1.00
[2]2[\mathbb{Q}_{2}]^{2} 6 2.92 ×105\times 10^{-5} 1.11 ×103\times 10^{-3}
12 3.90 ×106\times 10^{-6} 2.90 2.71 ×104\times 10^{-4} 2.04
24 5.02 ×107\times 10^{-7} 2.96 6.72 ×105\times 10^{-5} 2.01
48 6.36 ×108\times 10^{-8} 2.99 1.68 ×105\times 10^{-5} 2.00
[3]2[\mathbb{Q}_{3}]^{2} 6 3.02 ×107\times 10^{-7} 2.25 ×105\times 10^{-5}
12 2.17 ×108\times 10^{-8} 3.80 2.72 ×106\times 10^{-6} 3.05
24 1.45 ×109\times 10^{-9} 3.90 3.34 ×107\times 10^{-7} 3.03
48 9.33 ×1011\times 10^{-11} 3.96 4.13 ×108\times 10^{-8} 3.01
Table 1. Test 1: Convergence rates for 𝐐\mathbf{Q} with different degrees of polynomial approximation, in the decoupled case q=0q=0.

Regarding the density variation uu, we first present the convergence behaviour of the consistent discrete formulation eq. 13 with penalty parameter ϵ=1\epsilon=1, since we have proven the optimal error rate in the mesh-dependent norm ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}. The errors and convergence orders are listed in table 2. Optimal rates are observed in the ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}-norm. Furthermore, optimal orders of convergence in the L2L^{2}-norm are shown for approximating polynomials of degree greater than 22, while a sub-optimal rate in the L2L^{2}-norm is given for piecewise quadratic polynomials, exactly as expected. Sub-optimal convergence rates for quadratic polynomials were also illustrated in the numerical results of [29]. We also tested the convergence with the penalty parameter ϵ=5×104\epsilon=5\times 10^{4} and found that the discrete norms are very similar to table 2. We therefore avoid repeating the details here.

N=1hN=\frac{1}{h} eu0\|\textbf{e}_{u}\|_{0} rate eu1\|\textbf{e}_{u}\|_{1} rate |eu|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} rate
2\mathbb{Q}_{2} 6 1.17 ×105\times 10^{-5} 3.46 ×104\times 10^{-4} 1.36 ×102\times 10^{-2}
12 2.60 ×106\times 10^{-6} 2.17 9.81 ×105\times 10^{-5} 1.82 7.25 ×103\times 10^{-3} 0.91
24 6.37 ×107\times 10^{-7} 2.03 2.54 ×105\times 10^{-5} 1.95 3.54 ×103\times 10^{-3} 1.03
48 1.82 ×107\times 10^{-7} 1.80 6.88 ×106\times 10^{-6} 1.88 1.76 ×103\times 10^{-3} 1.01
3\mathbb{Q}_{3} 6 4.73 ×106\times 10^{-6} 1.32 ×104\times 10^{-4} 4.98 ×103\times 10^{-3}
12 3.32 ×107\times 10^{-7} 3.83 1.41 ×105\times 10^{-5} 3.23 9.96 ×104\times 10^{-4} 2.32
24 2.12 ×108\times 10^{-8} 3.97 1.63 ×106\times 10^{-6} 3.12 2.46 ×104\times 10^{-4} 2.02
48 1.32 ×109\times 10^{-9} 4.00 1.99 ×107\times 10^{-7} 3.03 6.14 ×105\times 10^{-5} 2.00
4\mathbb{Q}_{4} 6 2.01 ×107\times 10^{-7} 7.76 ×106\times 10^{-6} 3.94 ×104\times 10^{-4}
12 5.40 ×109\times 10^{-9} 5.22 4.30 ×107\times 10^{-7} 4.17 4.88 ×105\times 10^{-5} 3.01
24 1.68 ×1010\times 10^{-10} 5.00 2.68 ×108\times 10^{-8} 4.00 6.11 ×106\times 10^{-6} 2.99
48 5.27 ×1012\times 10^{-12} 4.99 1.68 ×109\times 10^{-9} 3.99 7.64 ×107\times 10^{-7} 3.00
Table 2. Test 1: Convergence rates using the consistent discrete formulation eq. 13 with penalty parameter ϵ=1\epsilon=1 and different polynomial degrees, in the decoupled case q=0q=0.

We next give the error rates for the inconsistent discrete formulation eq. 49. We illustrate the discrete norms and the computed convergence rates in table 3 with the penalty parameter ϵ=1\epsilon=1. It can be observed that only first order convergence is obtained in the H2H^{2}-like norm ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} even with different approximating polynomials. Moreover, we notice by comparing tables 2 and 3 that the convergence rate deteriorates slightly for polynomials of degree 3 (although not for degree 4). This, however, can be improved by choosing a larger penalty parameter, as shown in table 4 with ϵ=5×104\epsilon=5\times 10^{4}, where optimal rates are shown for the discrete norms ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h}, 1\|\cdot\|_{1} and 0\|\cdot\|_{0} for all polynomial degrees (except only sub-optimal in 0\|\cdot\|_{0} when a piecewise quadratic polynomial is used as the approximation). The inconsistent discrete formulation appears to be a reasonable choice when a sufficiently large penalty parameter is used.

N=1hN=\frac{1}{h} eu0\|\textbf{e}_{u}\|_{0} rate eu1\|\textbf{e}_{u}\|_{1} rate |eu|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} rate
2\mathbb{Q}_{2} 6 3.50 ×106\times 10^{-6} 1.06 ×104\times 10^{-4} 5.60 ×103\times 10^{-3}
12 8.76 ×108\times 10^{-8} 5.32 5.41 ×106\times 10^{-6} 4.29 2.56 ×103\times 10^{-3} 1.13
24 1.77 ×108\times 10^{-8} 2.31 7.47 ×107\times 10^{-7} 2.86 1.28 ×103\times 10^{-3} 0.99
48 4.35 ×109\times 10^{-9} 2.02 1.24 ×107\times 10^{-7} 2.56 6.42 ×104\times 10^{-4} 1.00
3\mathbb{Q}_{3} 6 6.47 ×106\times 10^{-6} 1.86 ×104\times 10^{-4} 7.59 ×103\times 10^{-3}
12 3.40 ×107\times 10^{-7} 4.25 1.73 ×105\times 10^{-5} 3.43 2.74 ×103\times 10^{-3} 1.47
24 1.98 ×108\times 10^{-8} 4.10 2.03 ×106\times 10^{-6} 3.09 1.31 ×103\times 10^{-3} 1.07
48 3.73 ×109\times 10^{-9} 2.39 2.63 ×107\times 10^{-7} 2.95 6.45 ×104\times 10^{-4} 1.02
4\mathbb{Q}_{4} 6 2.05 ×107\times 10^{-7} 7.85 ×106\times 10^{-6} 3.93 ×104\times 10^{-4}
12 5.40 ×109\times 10^{-9} 5.24 4.31 ×107\times 10^{-7} 4.19 4.88 ×105\times 10^{-5} 3.01
24 1.68 ×1010\times 10^{-10} 5.00 2.68 ×108\times 10^{-8} 4.01 6.11 ×106\times 10^{-6} 3.00
48 5.27 ×1012\times 10^{-12} 5.00 1.67 ×109\times 10^{-9} 4.00 7.64 ×107\times 10^{-7} 3.00
Table 3. Test 1: Convergence rates using the inconsistent discrete formulation eq. 49 with penalty parameter ϵ=1\epsilon=1 and different polynomial degrees, in the decoupled case q=0q=0.
N=1hN=\frac{1}{h} eu0\|\textbf{e}_{u}\|_{0} rate eu1\|\textbf{e}_{u}\|_{1} rate |eu|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} rate
2\mathbb{Q}_{2} 6 1.17 ×105\times 10^{-5} 3.48 ×104\times 10^{-4} 1.36 ×102\times 10^{-2}
12 2.62 ×106\times 10^{-6} 2.16 9.86 ×105\times 10^{-5} 1.82 7.26 ×103\times 10^{-3} 0.91
24 6.38 ×107\times 10^{-7} 2.04 2.54 ×105\times 10^{-5} 1.96 3.54 ×103\times 10^{-3} 1.03
48 1.82 ×107\times 10^{-7} 1.81 6.88 ×106\times 10^{-6} 1.88 1.76 ×103\times 10^{-3} 1.01
3\mathbb{Q}_{3} 6 4.80 ×106\times 10^{-6} 1.35 ×104\times 10^{-4} 4.92 ×103\times 10^{-3}
12 3.35 ×107\times 10^{-7} 3.84 1.43 ×105\times 10^{-5} 3.23 9.86 ×104\times 10^{-4} 2.32
24 2.14 ×108\times 10^{-8} 3.97 1.63 ×106\times 10^{-6} 3.13 2.45 ×104\times 10^{-4} 2.01
48 1.33 ×109\times 10^{-9} 4.01 1.99 ×107\times 10^{-7} 3.04 6.13 ×105\times 10^{-5} 2.00
4\mathbb{Q}_{4} 6 2.05 ×107\times 10^{-7} 7.85 ×106\times 10^{-6} 3.93 ×104\times 10^{-4}
12 5.40 ×109\times 10^{-9} 5.24 4.31 ×107\times 10^{-7} 4.19 4.88 ×105\times 10^{-5} 3.01
24 1.68 ×1010\times 10^{-10} 5.00 2.68 ×108\times 10^{-8} 4.01 6.11 ×106\times 10^{-6} 3.00
48 5.27 ×1012\times 10^{-12} 5.00 1.67 ×109\times 10^{-9} 4.00 7.64 ×107\times 10^{-7} 3.00
Table 4. Test 1: Convergence rates using the inconsistent discrete formulation eq. 49 with penalty parameter ϵ=5×104\epsilon=5\times 10^{4} and different polynomial degrees, in the decoupled case q=0q=0.

3.1.2. Convergence rate for q0q\neq 0

We next investigate the numerical convergence behaviour in the coupled case, i.e., q0q\neq 0, in this subsection. Its analysis remains future work, but since it is the coupled case that is solved in practice it is important to assure ourselves that the discretisation is sensible. For brevity, we fix the model parameter q=30q=30.

We examine the inconsistent discretisation for uu with penalty parameter ϵ=5×104\epsilon=5\times 10^{4}. In unreported preliminary experiments, we observed that the error in 𝐐\mathbf{Q} is governed by the lower of the degrees of the polynomials used for 𝐐\mathbf{Q} and uu. We thus give the convergence rates for uu and 𝐐\mathbf{Q} separately in tables 5 and 6, with the other degree fixed appropriately. It can be seen that 𝐐\mathbf{Q} retains optimal rates in both the H1H^{1} and L2L^{2} norms, and though there are some fluctuations of the order for uu, it still possesses very similar convergence rates when compared with the decoupled case described in table 4.

N=1hN=\frac{1}{h} eu0\|\textbf{e}_{u}\|_{0} rate eu1\|\textbf{e}_{u}\|_{1} rate |eu|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} rate
2\mathbb{Q}_{2} 6 1.21 ×105\times 10^{-5} 3.59 ×104\times 10^{-4} 1.37 ×102\times 10^{-2}
12 3.98 ×106\times 10^{-6} 1.61 1.42 ×104\times 10^{-4} 1.34 8.30 ×103\times 10^{-3} 0.72
24 1.57 ×106\times 10^{-6} 1.35 4.99 ×105\times 10^{-5} 1.51 3.89 ×103\times 10^{-3} 1.09
48 2.58 ×107\times 10^{-7} 2.60 9.06 ×106\times 10^{-6} 2.46 1.78 ×103\times 10^{-3} 1.13
3\mathbb{Q}_{3} 6 7.36 ×106\times 10^{-6} 2.25 ×104\times 10^{-4} 9.10 ×103\times 10^{-3}
12 4.13 ×107\times 10^{-7} 4.16 1.86 ×105\times 10^{-5} 3.60 1.11 ×103\times 10^{-3} 3.03
24 4.23 ×108\times 10^{-8} 3.29 2.24 ×106\times 10^{-6} 3.05 2.53 ×104\times 10^{-4} 2.14
48 3.01 ×109\times 10^{-9} 3.81 2.28 ×107\times 10^{-7} 3.29 6.15 ×105\times 10^{-5} 2.04
Table 5. Test 1: Convergence rates for uu with q=30q=30 and penalty parameter ϵ=5×104\epsilon=5\times 10^{4} with the inconsistent discretisation eq. 49 for uu, fixing the approximation for 𝐐\mathbf{Q} to be with the [2]2[\mathbb{Q}_{2}]^{2} element.
N=1hN=\frac{1}{h} e𝐐0\|\textbf{e}_{\mathbf{Q}}\|_{0} rate e𝐐1\|\textbf{e}_{\mathbf{Q}}\|_{1} rate
[1]2[\mathbb{Q}_{1}]^{2} 6 8.12 ×104\times 10^{-4} 3.78 ×102\times 10^{-2}
12 2.02 ×104\times 10^{-4} 2.01 1.88 ×102\times 10^{-2} 1.01
24 5.05 ×105\times 10^{-5} 2.00 9.39 ×103\times 10^{-3} 1.00
48 1.26 ×105\times 10^{-5} 2.00 4.69 ×103\times 10^{-3} 1.00
[2]2[\mathbb{Q}_{2}]^{2} 6 2.92 ×105\times 10^{-5} 1.11 ×103\times 10^{-3}
12 3.90 ×106\times 10^{-6} 2.90 2.71 ×104\times 10^{-4} 2.04
24 5.02 ×107\times 10^{-7} 2.96 6.72 ×105\times 10^{-5} 2.01
48 6.37 ×108\times 10^{-8} 2.98 1.68 ×105\times 10^{-5} 2.00
[3]2[\mathbb{Q}_{3}]^{2} 6 3.02 ×107\times 10^{-7} 2.25 ×105\times 10^{-5}
12 2.17 ×108\times 10^{-8} 3.80 2.72 ×106\times 10^{-6} 3.05
24 1.45 ×109\times 10^{-9} 3.90 3.34 ×107\times 10^{-7} 3.03
48 9.32 ×1011\times 10^{-11} 3.96 4.13 ×1008\times 10^{-08} 3.01
Table 6. Test 1: Convergence rates for 𝐐\mathbf{Q} with q=30q=30 and penalty parameter ϵ=5×104\epsilon=5\times 10^{4} with the inconsistent discretisation eq. 49 for uu, fixing the approximation for uu to be with the 3\mathbb{Q}_{3} element.
Remark 3.2.

We also tested the convergence with the consistent weak formulation for uu under the same numerical settings as in tables 5 and 6. We found that in both cases they present very similar convergence behaviour and thus we omit the details here.

3.2. Test 2: on the unit disc

For the second set of experiments, we provide numerical results for an exact solution ueu^{e} with only H3H^{3}-regularity, instead of CC^{\infty} as in eq. 62. Our goal is to investigate whether the H4H^{4}-regularity assumption on uu can be relaxed. To this end, we consider a triangular mesh of Ω={(x,y)|x2+y2<1}\Omega=\{(x,y)\ |\ x^{2}+y^{2}<1\} and choose ueu^{e} to be

ue=(x2+y2)3/2,u^{e}=(x^{2}+y^{2})^{3/2}, (63)

and choose the same exact solution for Q11eQ^{e}_{11} and Q12eQ^{e}_{12} as in eq. 62. The exact solution given by eq. 63 is in H3(Ω)H^{3}(\Omega) but not in H4(Ω)H^{4}(\Omega), hence violating the regularity assumption of the analysis in section 2.1.

The resulting convergence rates are reported in tables 7 and 8. Table 7 shows that optimal H1H^{1} and L2L^{2} rates are achieved for 𝐐\mathbf{Q} with three different choices of finite elements [1]2,[2]2,[3]2[\mathbb{P}_{1}]^{2},[\mathbb{P}_{2}]^{2},[\mathbb{P}_{3}]^{2}. Table 8 shows the convergence behaviour with penalty parameter ε=1\varepsilon=1 when using the inconsistent discrete formulation eq. 49. In contrast to table 3, only first order convergence is obtained for the discrete norm ||||||h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} and second-order convergence for 0\|\cdot\|_{0} and 1\|\cdot\|_{1}, with both 3\mathbb{P}_{3} and 4\mathbb{P}_{4}. Interestingly, table 8 indicates no convergence when using 2\mathbb{P}_{2} elements. It appears that the assumption uH4(Ω)u\in H^{4}(\Omega) is necessary for our analysis, and that a different analysis should be carried out when this assumption no longer holds.

No. of triangles e𝐐0\|\textbf{e}_{\mathbf{Q}}\|_{0} rate e𝐐1\|\textbf{e}_{\mathbf{Q}}\|_{1} rate
[1]2[\mathbb{P}_{1}]^{2} 60 6.08 ×102\times 10^{-2} 1.09
240 1.56 ×102\times 10^{-2} 1.96 5.80 ×101\times 10^{-1} 0.91
960 3.92 ×103\times 10^{-3} 2.00 2.93 ×101\times 10^{-1} 0.99
3840 9.83 ×104\times 10^{-4} 2.00 1.47 ×101\times 10^{-1} 1.00
15360 2.47 ×104\times 10^{-4} 1.99 7.34 ×102\times 10^{-2} 1.00
[2]2[\mathbb{P}_{2}]^{2} 60 8.97 ×103\times 10^{-3} 2.11 ×101\times 10^{-1}
240 1.51 ×103\times 10^{-3} 2.57 5.87 ×102\times 10^{-2} 1.84
960 2.22 ×104\times 10^{-4} 2.77 1.52 ×102\times 10^{-2} 1.95
3840 3.02 ×105\times 10^{-5} 2.88 3.85 ×103\times 10^{-3} 1.98
15360 3.93 ×106\times 10^{-6} 2.94 9.67 ×104\times 10^{-4} 1.99
[3]2[\mathbb{P}_{3}]^{2} 60 1.08 ×103\times 10^{-3} 3.21 ×102\times 10^{-2}
240 8.21 ×105\times 10^{-5} 3.72 4.58 ×103\times 10^{-3} 2.81
960 5.52 ×106\times 10^{-6} 3.89 5.92 ×104\times 10^{-4} 2.95
3840 3.54 ×107\times 10^{-7} 3.96 7.44 ×105\times 10^{-5} 2.99
15360 2.23 ×108\times 10^{-8} 3.99 9.31 ×106\times 10^{-6} 3.00
Table 7. Test 2: Convergence rates for 𝐐\mathbf{Q} with different degrees of polynomial approximation, in the decoupled case q=0q=0.
No. of triangles eu0\|\textbf{e}_{u}\|_{0} rate eu1\|\textbf{e}_{u}\|_{1} rate |eu|h{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\textbf{e}_{u}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{h} rate
2\mathbb{P}_{2} 60 1.36 ×102\times 10^{-2} 2.77 ×101\times 10^{-1} 7.39
240 1.58 ×103\times 10^{-3} 3.11 3.86 ×102\times 10^{-2} 2.84 2.39 1.63
960 7.76 ×104\times 10^{-4} 1.02 1.57 ×102\times 10^{-2} 1.30 1.44 0.73
3840 1.84 ×103\times 10^{-3} -1.25 3.42 ×102\times 10^{-2} -1.12 1.26 0.19
15360 2.77 ×103\times 10^{-3} -0.59 5.29 ×102\times 10^{-2} -0.63 1.60 -0.34
3\mathbb{P}_{3} 60 9.94 ×103\times 10^{-3} 2.21 ×101\times 10^{-1} 6.15
240 3.99 ×103\times 10^{-3} 1.32 8.49 ×102\times 10^{-2} 1.38 3.03 1.02
960 1.33 ×104\times 10^{-4} 4.90 4.57 ×103\times 10^{-3} 4.22 1.27 1.26
3840 3.21 ×105\times 10^{-5} 2.06 8.47 ×104\times 10^{-4} 2.43 0.66 0.93
15360 9.22 ×106\times 10^{-6} 1.80 2.11 ×104\times 10^{-4} 2.00 0.34 0.97
4\mathbb{P}_{4} 60 7.17 ×103\times 10^{-3} 1.65 ×101\times 10^{-1} 4.70
240 1.34 ×103\times 10^{-3} 2.42 3.84 ×102\times 10^{-2} 2.10 2.39 0.98
960 1.18 ×104\times 10^{-4} 3.50 4.54 ×103\times 10^{-3} 3.08 1.28 0.90
3840 2.98 ×105\times 10^{-5} 1.99 8.14 ×104\times 10^{-4} 2.48 0.67 0.94
15360 8.15 ×106\times 10^{-6} 1.87 1.86 ×104\times 10^{-4} 2.13 0.34 0.97
Table 8. Test 2: Convergence rates using the inconsistent discrete formulation eq. 49 with penalty parameter ϵ=1\epsilon=1 and different polynomial degrees, in the decoupled case q=0q=0.

References

  • [1] J. H. Argyris, I. Fried, and D. W. Scharpf, The TUBA family of plate elements for the matrix displacement method, Aeronaut. J., 72 (1968), pp. 701–709.
  • [2] J. M. Ball, Mathematics and liquid crystals, Mol. Cryst. Liq. Cryst., 647 (2017), pp. 1–27.
  • [3] J. M. Ball and S. J. Bedford, Discontinuous order parameters in liquid crystal theories, Mol. Cryst. Liq. Cryst., 612 (2015), pp. 1–23.
  • [4] S. J. Bedford, Calculus of variations and its application to liquid crystals, PhD thesis, University of Oxford, 2014.
  • [5] H. Blum and R. R. Bonn, On the boundary value problem of the biharmonic operator on domains with angular corners, Math. Mech. in the Appli. Sci., 2 (1980), pp. 556–581.
  • [6] J. P. Borthagaray, R. H. Nochetto, and S. W. Walker, A structure-preserving FEM for the uniaxially constrained Q-tensor model of nematic liquid crystals, Numer. Math., 145 (2020), pp. 837–881.
  • [7] S. C. Brenner, C0C^{0} interior penalty methods, in Frontiers in Numerical Analysis - Durham 2010. Lecture Notes in Computational Science and Engineering, J. Blowey and M. Jensen, eds., vol. 85, Springer, Berlin, Heidelberg, 2011.
  • [8] S. C. Brenner, T. Gudi, and L. Sung, A weakly over-penalized symmetric interior penalty method for the biharmonic problem, Electon. Trans. Numer. Anal., 37 (2010), pp. 214–238.
  • [9] S. C. Brenner and L. Sung, C0C^{0} interior penalty methods for fourth order elliptic boundary value problems on polygonal domains, J. Sci. Comput., 22 (2005), pp. 83–118.
  • [10] S. C. Brenner and L. Sung, A weakly over-penalized symmetric interior penalty method, Electon. Trans. Numer. Anal., 30 (2008), pp. 107–127.
  • [11] S. C. Brenner, K. Wang, and J. Zhao, Poincaré-Friedrichs inequalities for piecewise H2H^{2} functions, Numer. Funct. Anal. Optim., 25 (2004), pp. 463–478.
  • [12] X. Cheng, W. Han, and H. Huang, Some mixed finite element methods for the biharmonic equation, J. Comp. Appl. Math., 126 (2000), pp. 91–109.
  • [13] T. A. Davis, Finite element analysis of the Landau–de Gennes minimization problem for liquid crystals in confinement, PhD thesis, Kent State University, 1994.
  • [14] T. A. Davis and J. E. C. Gartland, Finite element analysis of the Landau-de Gennes minimization problem for liquid crystals, SIAM J. Numer. Anal., 35 (1998), pp. 336–362.
  • [15] P. G. de Gennes, An analogy between superconductors and smectic A, Solid State Commun., 10 (1972), pp. 753–756.
  • [16] P. G. de Gennes, The Physics of Liquid Crystals, Oxford University Press, Oxford, 1974.
  • [17] G. Engel, K. Garikipati, T. J. R. Hughes, M. G. Larson, L. Mazzei, and R. L. Taylor, Continuous/discontinuous finite element approximations of fourth-order elliptic problems in structural and continuum mechanics with applications to thin beams and plates, and strain gradient elasticity, Comput. Methods Appl. Mech. Engrg., 191 (2002), pp. 3669–3750.
  • [18] L. C. Evans, Partial Differential Equations, vol. 19 of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, 2nd ed., 2010.
  • [19] C. J. García-Cervera and S. Joo, Layer undulations in Smectic A liquid crystals, J. Comp. Theo. Nano., 7 (2010), pp. 795–801.
  • [20] S. Kesavan, Topics in Functional Analysis and Applications, John Wiley & Sons, New York, 1989.
  • [21] R. R. Maity, A. Majumdar, and N. Nataraj, Discontinuous Galerkin finite element methods for the Landau–de Gennes minimization problem of liquid crystals, IMA J. Numer. Anal., 41 (2021), pp. 1130–1163.
  • [22] P. Monderkamp, R. Wittmann, L. B. G. Cortes, D. G. A. L. Aarts, and H. Löwen, Topology of orientational defects in confined smectic liquid crystals, Phys. Rev. Lett., 127 (2021), pp. 1–10.
  • [23] N. J. Mottram and C. J. P. Newton, Introduction to 𝐐\mathbf{Q}-tensor theory, 2014, https://arxiv.org/abs/1409.3542.
  • [24] M. Y. Pevnyi, J. Selinger, and T. J. Sluckin, Modeling smectic layers in confined geometries: order parameter and defects, Phys. Rev. E, 90 (2014), pp. 1–8.
  • [25] M. Robinson, C. Luo, P. E. Farrell, R. Erban, and A. Majumdar, From molecular to continuum modelling of bistable liquid crystal devices, Liq. Cryst., 44 (2017), pp. 2267–2284.
  • [26] H. Sackmann, Plenary lecture. Smectic liquid crystals. A historical review, Liq. Cryst., 5 (1989), pp. 43–55.
  • [27] R. Scholtz, A mixed method for fourth-order problems using linear finite elements, RAIRO Numer. Anal., 15 (1978), pp. 85–90.
  • [28] I. W. Stewart, The Static and Dynamic Continuum Theory of Liquid Crystals: A Mathematical Introduction, CPC Press, 2004.
  • [29] E. Süli and I. Mozolevski, hphp-version interior penalty DGFEMs for the biharmonic equation, Comput. Methods Appl. Mech. Engrg., 196 (2007), pp. 1851–1863.
  • [30] R. Wittmann, L. B. G. Cortes, H. Löwen, and D. G. A. L. Aarts, Particle-resolved topological defects of smectic colloidal liquid crystals in extreme confinement, Nat. Comm., 12 (2021), pp. 1–10.
  • [31] J. Xia, S. MacLachlan, T. J. Atherton, and P. E. Farrell, Structural landscapes in geometrically frustrated smectics, Phys. Rev. Lett., 126 (2021), pp. 1–6.
  • [32] S. Zhang, A family of 3D continuously differentiable finite elements on tetrahedral grids, Appl. Numer. Math., 59 (2009), pp. 219–233.