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

Mixed finite element methods for the ferrofluid model with magnetization paralleled to the magnetic field

Yongke Wu and Xiaoping Xie
Abstract.

Mixed finite element methods are considered for a ferrofluid flow model with magnetization paralleled to the magnetic field. The ferrofluid model is a coupled system of the Maxwell equations and the incompressible Navier-Stokes equations. By skillfully introducing some new variables, the model is rewritten as several decoupled subsystems that can be solved independently. Mixed finite element formulations are given to discretize the decoupled systems with proper finite element spaces. Existence and uniqueness of the mixed finite element solutions are shown, and optimal order error estimates are obtained under some reasonable assumptions. Numerical experiments confirm the theoretical results.

Key words and phrases:
ferrofluid flow, decoupled system, mixed finite element method, error estimate
2010 Mathematics Subject Classification:
65N55; 65F10; 65N22; 65N30;
Y. Wu was supported by the National Natural Science Foundation of China (11971094 and 11501088), and X. Xie was supported by the National Nature Science Foundation of China (12171340)
Y. Wu, School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu 611731, China. Email: [email protected]
X. Xie (Corresponding author), School of Mathematical, Sichuan University, Chengdu 610064, China. Email: [email protected]

1. Introduction

Ferrofluids are colloidal liquids consisting of non-conductive nanoscale ferromagnetic or ferrimagnetic particles suspended in carrier fluids, and have extensive applications in many technology and biomedicine fields  [22, 31]. There are two main ferrohydrodynamics (FHD) models which treat ferrofluids as homogeneous monophase fluids: the Rosensweig’s model  [24, 23] and the Shliomis’ model [25, 26]. The main difference between these two models lies in that the former one considers the internal rotation of the nanoparticles, while the latter deals with the rotation as a magnetic torque. We refer to [1, 4, 2, 3, 21] for some existence results of solutions to the two FHD models.

The FHD models are coupled nonlinear systems of the Maxwell equations and the incompressible Navier-Stokes equations. There are limited works in the literature on the numerical analysis of the FHD models. In [27, 17, 16, 30], several numerical schemes were applied to solve reduced two dimensional FHD models where some nonlinear terms of the original models are dropped. Nochetto et al. [20] showed the energy stability of the Rosensweig’s model, proposed an energy-stable numerical scheme using finite element methods, and gave the existence and convergence of the numerical solutions. Recently, Wu and Xie [29] developed a class of energy-preserving mixed finite element methods for the Shliomis’ FHD model, and derived optimal error estimates for both the the semi- and fully discrete schemes. We also note that in [32] an unconditionally energy-stable fully discrete finite element method was presented for a two-phase FHD model.

In this paper, we consider the Shliomis’ FHD model with the assumption that the magnetization field is parallel to the magnetic field. Under this assumption, the magnetization equation in the Shliomis’ model degenerates to the Langevin magnetization law [17, 24, 23]. We introduce some new variables to transform the model into two main decoupled subsystems, i.e., a nonlinear elliptic equation and the incompressible Navier-Stokes equations. We apply proper finite element spaces to discretize the nonlinear decoupled systems, prove the existence and, under some reasonable assumptions, uniqueness of the finite element solutions, and derive optimal error estimates. We also show that our scheme preserves the ferrofluids’ nonconductive property curl𝑯=0{\rm curl\,}\boldsymbol{H}=0 exactly.

The rest of this paper is organized as follows. In section 2, we introduce several Sobolev spaces, give the governing equation of the FHD model with magnetization paralleled to magnetic field, reform the FHD model equivalently, and construct the weak formulations. In section 3, we recall the finite element spaces, show the existence and uniqueness of solutions for the constructed finite element methods, and give the optimal order error estimates. In section 4, some numerical experiments will be given to verify our theoretical results.

2. Preliminary

In this section, we introduce several Sobolev spaces, give the governing equations of the FHD model with magnetization paralleled to magnetic field, derive the equivalent decoupled systems, and present the weak formulations.

2.1. Sobolev spaces

Let Ωd\Omega\subset\mathbb{R}^{d} (d=2,3d=2,3) be a bounded and simply connected convex domain with Lipschitz boundary Ω\partial\Omega, and 𝒏\boldsymbol{n} be the unit outward normal vector on Ω\partial\Omega.

For any p1p\geq 1, we denote by Lp(Ω)L^{p}(\Omega) the space of all power-pp integrable functions on Ω\Omega with norm Lp\|\cdot\|_{L^{p}}. For any nonnegative integer mm, denote by Hm(Ω)H^{m}(\Omega) the usual mm-th order Sobolev space with norm m\|\cdot\|_{m} and semi-norm ||m|\cdot|_{m}. In particular, H0(Ω)=L2(Ω)H^{0}(\Omega)=L^{2}(\Omega) denotes the space of square integrable functions on Ω\Omega, with the inner product (,)(\cdot,\cdot) and norm \|\cdot\|. For the vector spaces 𝑯m(Ω):=(Hm(Ω))d\boldsymbol{H}^{m}(\Omega):=(H^{m}(\Omega))^{d} and 𝑳2(Ω):=(L2(Ω))d\boldsymbol{L}^{2}(\Omega):=(L^{2}(\Omega))^{d}, we use the same notations of norm, semi-norm and inner product as those for the scalar cases.

We further introduce the Sobolev spaces

𝑯(curl):={{𝒗(L2(Ω))2:curl𝒗L2(Ω)}if d=2,{𝒗(L2(Ω))3:curl𝒗(L2(Ω))3}if d=3\boldsymbol{H}({\rm curl\,}):=\left\{\begin{array}[]{ll}\{\boldsymbol{v}\in(L^{2}(\Omega))^{2}:~{}~{}{\rm curl\,}\boldsymbol{v}\in L^{2}(\Omega)\}&\text{if }d=2,\\ \{\boldsymbol{v}\in(L^{2}(\Omega))^{3}:~{}~{}{\rm curl\,}\boldsymbol{v}\in(L^{2}(\Omega))^{3}\}&\text{if }d=3\end{array}\right.

and

𝑯(div):={𝒗(L2(Ω))d:div𝒗L2(Ω)},\boldsymbol{H}(\operatorname{div}):=\{\boldsymbol{v}\in(L^{2}(\Omega))^{d}:~{}~{}\operatorname{div}\boldsymbol{v}\in L^{2}(\Omega)\},

and set

H01(Ω):={vH1(Ω):v=0 on Ω},\displaystyle H_{0}^{1}(\Omega):=\{v\in H^{1}(\Omega):~{}v=0\text{ on }\partial\Omega\},
𝑯0(curl):={𝒗𝑯(curl):𝒗×𝒏=0 on Ω},\displaystyle\boldsymbol{H}_{0}({\rm curl\,}):=\{\boldsymbol{v}\in\boldsymbol{H}({\rm curl\,}):~{}\boldsymbol{v}\times\boldsymbol{n}=0\text{ on }\partial\Omega\},
𝑯0(div):={𝒗𝑯(div):𝒗𝒏=0 on Ω},\displaystyle\boldsymbol{H}_{0}(\operatorname{div}):=\{\boldsymbol{v}\in\boldsymbol{H}(\operatorname{div}):~{}\boldsymbol{v}\cdot\boldsymbol{n}=0\text{ on }\partial\Omega\},
L02(Ω):={vL2(Ω):ΩvdΩ=0},\displaystyle L_{0}^{2}(\Omega):=\{v\in L^{2}(\Omega):~{}\int_{\Omega}v\,{\rm d}\Omega=0\},

where

curl𝒗:={xv2yv1if 𝒗=(v1,v2),(yv3zv2,zv1xv3,xv2yv1)if 𝒗=(v1,v2,v3).\displaystyle{\rm curl\,}\boldsymbol{v}:=\left\{\begin{array}[]{ll}\partial_{x}v_{2}-\partial_{y}v_{1}&\text{if }\boldsymbol{v}=(v_{1},~{}v_{2})^{\intercal},\\ (\partial_{y}v_{3}-\partial_{z}v_{2},~{}\partial_{z}v_{1}-\partial_{x}v_{3},~{}\partial_{x}v_{2}-\partial_{y}v_{1})^{\intercal}&\text{if }\boldsymbol{v}=(v_{1},~{}v_{2},~{}v_{3})^{\intercal}.\end{array}\right.
div𝒗:={xv1+yv2if 𝒗=(v1,v2),xv1+yv2+zv3if 𝒗=(v1,v2,v3),\displaystyle\operatorname{div}\boldsymbol{v}:=\left\{\begin{array}[]{ll}\partial_{x}v_{1}+\partial_{y}v_{2}&\text{if }\boldsymbol{v}=(v_{1},v_{2})^{\intercal},\\ \partial_{x}v_{1}+\partial_{y}v_{2}+\partial_{z}v_{3}&\text{if }\boldsymbol{v}=(v_{1},v_{2},v_{3})^{\intercal},\end{array}\right.

and (x,y)(x,y) and (x,y,z)(x,y,z) are the Cartesian coordinates in two and three dimensions, respectively.

For any Sobolev space SS with norm S\|\cdot\|_{S}, we use SS^{\prime} to denote the dual space of SS, and ,\langle\cdot,\cdot\rangle to denote the dual product between SS^{\prime} and SS. For any fSf\in S^{\prime}, the operator norm of ff is defined as fS=sup0vSf,vvS\|f\|_{S^{\prime}}=\sup\limits_{0\neq v\in S}\frac{\langle f,v\rangle}{\|v\|_{S}}.

2.2. Governing equations of the ferrofluid flow

We consider the domain Ω\Omega filled with ferrofluid flow. On the macroscopic level, the mathematical model for describing the interactions between magnetic fields and ferrofluids consists of the Maxwell’s equations and the Navier-Stokes equations [23, 24, 25, 26]. Since the ferrofluid flow is nonconductive, the corresponding Maxwell’s equations read as:

(1) curl𝑯=0inΩ,\displaystyle{\rm curl\,}\boldsymbol{H}=0\qquad\qquad\qquad\,\text{in}\quad\Omega,
(2) div𝑩=div𝑯einΩ,\displaystyle\operatorname{div}\boldsymbol{B}=\operatorname{div}\boldsymbol{H}_{e}\qquad\qquad\text{in}\quad\Omega,

with the magnetic field 𝑯\boldsymbol{H} and the magnetic induction 𝑩\boldsymbol{B} satisfying the relation

(3) 𝑩=μ0(𝑯+𝑴)inΩ,\boldsymbol{B}=\mu_{0}(\boldsymbol{H}+\boldsymbol{M})\qquad\qquad\text{in}\quad\Omega,

where μ0>0\mu_{0}>0 is the permeability constant, 𝑴\boldsymbol{M} is the magnetization, and 𝑯e\boldsymbol{H}_{e} is the known external magnetic field that satisfies 𝑯e𝒏=0\boldsymbol{H}_{e}\cdot\boldsymbol{n}=0 on Ω\partial\Omega.

Under the assumption that the magnetization 𝑴\boldsymbol{M} of the ferrofluid flow is parallel to the magnetic field 𝑯\boldsymbol{H}, it follows the nonlinear Langevin magnetization law [23, 24, 17], i.e.,

(4) 𝑴(𝑯)=Ms(coth(γH)1γH)𝑯H,\boldsymbol{M}(\boldsymbol{H})=M_{s}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{\boldsymbol{H}}{H},

with the saturation magnetization Ms>0M_{s}>0, the Lagevin parameter γ=3χ0/Ms\gamma=3\chi_{0}/M_{s}, the initial susceptibility χ0>0\chi_{0}>0, H:=|𝑯|=(𝑯𝑯)1/2>0H:=|\boldsymbol{H}|=(\boldsymbol{H}\cdot\boldsymbol{H})^{1/2}>0, and coth(x)=ex+exexex\coth(x)=\frac{e^{x}+e^{-x}}{e^{x}-e^{-x}}.

The hydrodynamic properties of the ferrofluid flow are described by the incompressible Navier-Stokes equations

(5) ρ(𝒖)𝒖ηΔ𝒖+p=𝒇+μ0(𝑴)𝑯inΩ,\displaystyle\rho(\boldsymbol{u}\cdot\nabla)\boldsymbol{u}-\eta\Delta\boldsymbol{u}+\nabla p=\boldsymbol{f}+\mu_{0}(\boldsymbol{M}\cdot\nabla)\boldsymbol{H}\qquad\quad\text{in}\quad\Omega,
(6) div𝒖=0inΩ,\displaystyle\operatorname{div}\boldsymbol{u}=0\,\,\,\,\,\quad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\text{in}\quad\Omega,

where ρ\rho denotes the fluid density, 𝒖\boldsymbol{u} the velocity field of the flow, η\eta the dynamic viscosity, and 𝒇\boldsymbol{f} the known volume force.

We consider the following homogenous boundary conditions for equation (1)-(6):

(7) 𝒖=0 and 𝑯×𝒏=0onΩ.\boldsymbol{u}=0\quad\text{ and }\quad\boldsymbol{H}\times\boldsymbol{n}=0\qquad\text{on}\quad\partial\Omega.

Using the fact that

(𝑯)𝑯=12H2𝑯×(curl𝑯)=12H2(\boldsymbol{H}\cdot\nabla)\boldsymbol{H}=\frac{1}{2}\nabla H^{2}-\boldsymbol{H}\times({\rm curl\,}\boldsymbol{H})=\frac{1}{2}\nabla H^{2}

and the Langevin magnetization law (4), we have

(8) (𝑴)𝑯=MsH(coth(γH)1γH)12H2=Msγlnsinh(γH)H,(\boldsymbol{M}\cdot\nabla)\boldsymbol{H}=\frac{M_{s}}{H}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{1}{2}\nabla H^{2}=\frac{M_{s}}{\gamma}\nabla\ln\frac{\sinh(\gamma H)}{H},

where sinh(x)=exex2\sinh(x)=\frac{e^{x}-e^{-x}}{2}. Denote

(9) β(x):=Msγlnsinh(γx)x,\beta(x):=\frac{M_{s}}{\gamma}\ln\frac{\sinh(\gamma x)}{x},

and introduce two variables

(10) ψ:=β(H)\psi:=\beta(H)

and

(11) p~:=pμ0ψ,\tilde{p}:=p-\mu_{0}\psi,

then equation (5) becomes

(12) ρ(𝒖)𝒖ηΔ𝒖+p~=𝒇in Ω.\rho(\boldsymbol{u}\cdot\nabla)\boldsymbol{u}-\eta\Delta\boldsymbol{u}+\nabla\tilde{p}=\boldsymbol{f}\qquad\qquad\qquad\text{in }\Omega.

Equation (1) and the boundary condition 𝑯×𝒏|Ω=0\boldsymbol{H}\times\boldsymbol{n}|_{\partial\Omega}=0 in (7) imply that (cf. [6]) there exists ϕH01(Ω)\phi\in H_{0}^{1}(\Omega) with

(13) 𝑯=ϕ,\boldsymbol{H}=\nabla\phi,

then combining (2), (3) and (4) leads to

(14) (α(|ϕ|)ϕ)=1μ0div𝑯e=:ginΩ,\nabla\cdot\left(\alpha(|\nabla\phi|)\nabla\phi\right)=\frac{1}{\mu_{0}}\operatorname{div}\boldsymbol{H}_{e}=:g\qquad\text{in}~{}~{}\Omega,

with

(15) α(x):=1+Msx(coth(γx)1γx).\alpha(x):=1+\frac{M_{s}}{x}\left(\coth(\gamma x)-\frac{1}{\gamma x}\right).

The above equivalence transformation shows that the FHD model (1) - (6) with the boundary conditions (7) can be equivalently written as follows: Find ϕH01(Ω)\phi\in H_{0}^{1}(\Omega), 𝑯𝑯0(curl)\boldsymbol{H}\in\boldsymbol{H}_{0}({\rm curl\,}), 𝑴𝑯0(curl)\boldsymbol{M}\in\boldsymbol{H}_{0}({\rm curl\,}), 𝒖𝑯01(Ω)\boldsymbol{u}\in\boldsymbol{H}_{0}^{1}(\Omega), p~L02(Ω)\tilde{p}\in L_{0}^{2}(\Omega), ψL02(Ω)\psi\in L_{0}^{2}(\Omega), and pL02(Ω)p\in L_{0}^{2}(\Omega) such that

(16) {(α(|ϕ|)ϕ)=gin Ω,𝑯ϕ=0in Ω,𝑴=Ms(coth(γH)1γH)𝑯H=(α(H)1)𝑯in Ω,ρ(𝒖)𝒖ηΔ𝒖+p~=𝒇in Ω,𝒖=0in Ω,ψ=Msγlnsinh(γH)H=β(H)in Ω,p=p~+μ0ψin Ω.\left\{\begin{array}[]{ll}\nabla\cdot(\alpha(|\nabla\phi|)\nabla\phi)=g&\text{in }\Omega,\\ \boldsymbol{H}-\nabla\phi=0&\text{in }\Omega,\\ \boldsymbol{M}=M_{s}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{\boldsymbol{H}}{H}=\big{(}\alpha(H)-1\big{)}\boldsymbol{H}&\text{in }\Omega,\\ \rho(\boldsymbol{u}\cdot\nabla)\boldsymbol{u}-\eta\Delta\boldsymbol{u}+\nabla\tilde{p}=\boldsymbol{f}&\text{in }\Omega,\\ \nabla\cdot\boldsymbol{u}=0&\text{in }\Omega,\\ \psi=\frac{M_{s}}{\gamma}\ln\frac{\sinh(\gamma H)}{H}=\beta(H)&\text{in }\Omega,\\ p=\tilde{p}+\mu_{0}\psi&\text{in }\Omega.\end{array}\right.
Remark 2.1.

The system (16) is a decoupled system. Firstly, we can solve the first nonlinear elliptic equation of (16) to get ϕ\phi, and solve the Navier-Stokes equations, i.e. the fourth and fifth equations, to get 𝐮\boldsymbol{u} and p~\tilde{p}. Secondly, we can get 𝐇\boldsymbol{H} from the second equation of (16). Finally, we can obtain 𝐌\boldsymbol{M}, ψ\psi, and pp from the third, the sixth, and the seventh equations, respectively.

2.3. Preliminary estimates for nonlinear functions

Notice that the decoupled system (16) involves the nonlinear functions α(x)\alpha(x) and β(x)\beta(x) defined in (15) and (9), respectively. The following basic estimates of these two functions will be used in later analysis.

Lemma 2.2.
  • (i)

    There exist a positive constants C1C_{1} and CαC_{\alpha} such that for any x>0x>0, there hold

    1<α(x)C1,|α(x)|Cα;1<\alpha(x)\leq C_{1},\qquad|\alpha^{\prime}(x)|\leq C_{\alpha}\ ;
  • (ii)

    For any x>0x>0, there holds

    |β(x)|Ms.|\beta^{\prime}(x)|\leq M_{s}.
Proof.

(i) We first show 1<α(x)C1.1<\alpha(x)\leq C_{1}. On one hand, the L’Hopital law implies that

limy0+ycoth(y)=1,\lim\limits_{y\rightarrow 0^{+}}y\coth(y)=1,

which, together with the fact that

(ycoth(y))=coth(y)ycsch2(y)=e2ye2y4y(eyey)2>0y>0,(y\coth(y))^{\prime}=\coth(y)-y{\rm csch}^{2}(y)=\frac{e^{2y}-e^{-2y}-4y}{(e^{y}-e^{-y})^{2}}>0\quad\forall~{}~{}y>0,

shows

ycoth(y)>1y>0.y\coth(y)>1\qquad\forall~{}~{}y>0.

Therefore,

α(x)=1+γMsycoth(y)1y2>1with y=γx, for x>0.\alpha(x)=1+\gamma M_{s}\frac{y\coth(y)-1}{y^{2}}>1\qquad\text{with }y=\gamma x,\text{ for }\forall~{}~{}x>0.

On the other hand, we easily see that

limx0+α(x)=1+γMs3andlimx+α(x)=1,\lim\limits_{x\rightarrow 0^{+}}\alpha(x)=1+\frac{\gamma M_{s}}{3}\qquad\text{and}\qquad\lim\limits_{x\rightarrow+\infty}\alpha(x)=1,

which, together the continuity of α(x)\alpha(x) for x>0x>0, indicate that there exists a positive constant C1C_{1} such that

α(x)C1x>0.\alpha(x)\leq C_{1}\qquad\forall~{}~{}x>0.

Second, let us show |α(x)|Cα.|\alpha^{\prime}(x)|\leq C_{\alpha}. It is easy to get

α(x)=2Msγx3Msx2coth(γx)γMsxcsch2(γx)=γ2Msy3(2y2csch2(y)ycoth(y))with y=γx.\begin{split}\alpha^{\prime}(x)&=\frac{2M_{s}}{\gamma x^{3}}-\frac{M_{s}}{x^{2}}\coth(\gamma x)-\frac{\gamma M_{s}}{x}{\rm csch}^{2}(\gamma x)\\ &=\frac{\gamma^{2}M_{s}}{y^{3}}(2-y^{2}{\rm csch}^{2}(y)-y\coth(y))\qquad\qquad\text{with }y=\gamma x.\end{split}

The L’Hopital law implies that

limx0+α(x)=0andlimx+α(x)=0.\lim\limits_{x\rightarrow 0^{+}}\alpha^{\prime}(x)=0\qquad\text{and}\qquad\lim\limits_{x\rightarrow+\infty}\alpha^{\prime}(x)=0.

Since α(x)\alpha^{\prime}(x) is continuous for x>0x>0, we know that there exists a positive constant CαC_{\alpha} such that

|α(x)|Cαx>0.|\alpha^{\prime}(x)|\leq C_{\alpha}\qquad\forall x>0.

As a result, the conclusions of (i) follow.

(ii) It is easy to find that

β(x)=Ms(coth(γx)1γx)=Ms(coth(y)1y)\beta^{\prime}(x)=M_{s}\left(\coth(\gamma x)-\frac{1}{\gamma x}\right)=M_{s}\left(\coth(y)-\frac{1}{y}\right)

with y=γx>0y=\gamma x>0. The L’Hopital law implies that

limx0+β(x)=0andlimx+β(x)=Ms.\lim\limits_{x\rightarrow 0^{+}}\beta^{\prime}(x)=0\quad\text{and}\quad\lim\limits_{x\rightarrow+\infty}\beta^{\prime}(x)=M_{s}.

The fact that

(coth(y)1y)=1y2csch2(y)y2=1+ycsch(y)y2eyey2yeyey>0y>0\left(\coth(y)-\frac{1}{y}\right)^{\prime}=\frac{1-y^{2}{\rm csch}^{2}(y)}{y^{2}}=\frac{1+y{\rm csch}(y)}{y^{2}}\frac{e^{y}-e^{-y}-2y}{e^{y}-e^{-y}}>0\qquad\forall~{}~{}y>0

implies that β(x)\beta^{\prime}(x) is a monotonically increasing function on the interval (0,+)(0,+\infty). We conclude that

0<β(x)Ms.0<\beta^{\prime}(x)\leq M_{s}.

This finishes the proof. ∎

2.4. Weak formulations

Based on the FHD model (16) and Remark 2.1, we consider the following weak formulations: Find ϕH01(Ω)\phi\in H_{0}^{1}(\Omega), 𝑯𝑯0(curl)\boldsymbol{H}\in\boldsymbol{H}_{0}({\rm curl\,}), 𝒖𝑯01(Ω)\boldsymbol{u}\in\boldsymbol{H}_{0}^{1}(\Omega) and p~L02(Ω)\tilde{p}\in L_{0}^{2}(\Omega) such that

(17) a(ϕ;ϕ,τ)=(g,τ)τH01(Ω),\displaystyle a(\phi;\phi,\tau)=-(g,\tau)\qquad\qquad\qquad\qquad\qquad\qquad\qquad\ \,\forall~{}~{}\tau\in H_{0}^{1}(\Omega),
(18) (𝑯,𝑪)(ϕ,𝑪)=0𝑪𝑯0(curl),\displaystyle(\boldsymbol{H},\boldsymbol{C})-(\nabla\phi,\boldsymbol{C})=0\quad\qquad\qquad\qquad\qquad\qquad\qquad\forall~{}~{}\boldsymbol{C}\in\boldsymbol{H}_{0}({\rm curl\,}),
(19) (ρ(𝒖)𝒖,𝒗)+η(𝒖,𝒗)(p~,𝒗)=(𝒇,𝒗)𝒗𝑯01(Ω),\displaystyle(\rho(\boldsymbol{u}\cdot\nabla)\boldsymbol{u},\boldsymbol{v})+\eta(\nabla\boldsymbol{u},\nabla\boldsymbol{v})-(\tilde{p},\nabla\cdot\boldsymbol{v})=(\boldsymbol{f},\boldsymbol{v})\qquad\forall~{}~{}\boldsymbol{v}\in\boldsymbol{H}_{0}^{1}(\Omega),
(20) (𝒖,q)=0qL02(Ω),\displaystyle(\nabla\cdot\boldsymbol{u},q)=0\,\quad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\forall~{}~{}q\in L_{0}^{2}(\Omega),

where a(;,):H1(Ω)×H1(Ω)×H1(Ω)a(\cdot;\cdot,\cdot):~{}H^{1}(\Omega)\times H^{1}(\Omega)\times H^{1}(\Omega)\rightarrow\mathbb{R} is defined by

a(w;ϕ,τ):=Ωα(|w|)ϕτdΩw,ϕ,τH1(Ω).a(w;\phi,\tau):=\int_{\Omega}\alpha(|\nabla w|)\nabla\phi\cdot\nabla\tau\,{\rm d}\Omega\qquad\forall~{}~{}w,~{}\phi,~{}\tau\in H^{1}(\Omega).
Remark 2.3.

As shown in Remark 2.1, once the variables ϕ\phi, 𝐇\boldsymbol{H}, 𝐮\boldsymbol{u} and p~\tilde{p} are solved, the other variables, i.e. 𝐌\boldsymbol{M}, ψ\psi and pp, can immediately be obtained by the third, the sixth and the seventh equations of (16), respectively.

In what follows, we discuss the existence and uniqueness of the solutions to the weak formulations (17) - (20).

We first consider the nonlinear equation (17). It is easy to see that, for any given wH1(Ω)w\in H^{1}(\Omega), a(w;,)a(w;\cdot,\cdot) is a bilinear form on H1(Ω)H^{1}(\Omega). Recall that the Poincaré inequality

vCpvvH01(Ω),\|v\|\leq C_{p}\|\nabla v\|\qquad\forall~{}~{}v\in H_{0}^{1}(\Omega),

with Cp>0C_{p}>0 being a constant depending only on Ω\Omega, means that the semi-norm ()\|\nabla(\cdot)\| is also a norm on H01(Ω)H_{0}^{1}(\Omega). Then, from Lemma 2.2 (i), we easily obtain the following uniform coercivity and continuity results for a(w;,)a(w;\cdot,\cdot):

Lemma 2.4.

For any wH01(Ω)w\in H_{0}^{1}(\Omega) and ϕ,τH01(Ω)\phi,\tau\in H_{0}^{1}(\Omega), we have

a(w;ϕ,ϕ)ϕ2a(w;\phi,\phi)\geq\|\nabla\phi\|^{2}

and

a(w;ϕ,τ)C1ϕτ.a(w;\phi,\tau)\leq C_{1}\|\nabla\phi\|\|\nabla\tau\|.

We have the following wellposedness results for equation (17).

Theorem 2.5.

The nonlinear equation (17) has at least one solution ϕH01(Ω)\phi\in H_{0}^{1}(\Omega), and there holds

(21) ϕ1μ0𝑯e.\|\nabla\phi\|\leq\frac{1}{\mu_{0}}\|\boldsymbol{H}_{e}\|.

Moreover, (17) admits at most one solution ϕH01(Ω)\phi\in H_{0}^{1}(\Omega) satisfying

(22) ϕL<1/Cα.\|\nabla\phi\|_{L^{\infty}}<1/C_{\alpha}.
Proof.

The existence of a solution ϕH01(Ω)\phi\in H_{0}^{1}(\Omega) follows from Lemma 2.4 and [8, Page 332, Theorem 2] directly.

Recalling that g=1μ0div𝑯eg=\frac{1}{\mu_{0}}\operatorname{div}\boldsymbol{H}_{e}, from (17) and the above lemma we have

ϕ2a(ϕ;ϕ,ϕ)=(g,ϕ)=1μ0(𝑯e,ϕ)1μ0𝑯eϕ.\|\nabla\phi\|^{2}\leq a(\phi;\phi,\phi)=-(g,\phi)=\frac{1}{\mu_{0}}(\boldsymbol{H}_{e},\nabla\phi)\leq\frac{1}{\mu_{0}}\|\boldsymbol{H}_{e}\|\|\nabla\phi\|.

This yields the estimate (21).

The thing left is to show the uniqueness of the solution under condition (22). In fact, let ϕ\phi and ϕ~\tilde{\phi} be any two solutions of (17) satisfying (22), we have

a(ϕ~;ϕ~,τ)=a(ϕ;ϕ,τ)τH01(Ω),a(\tilde{\phi};\tilde{\phi},\tau)=a(\phi;\phi,\tau)\qquad\forall~{}~{}\tau\in H_{0}^{1}(\Omega),

which yields

a(ϕ~;ϕ~ϕ,τ)=Ω(α(|ϕ|)α(|ϕ~|))ϕτdΩτH01(Ω).\displaystyle a(\tilde{\phi};\tilde{\phi}-\phi,\tau)=\int_{\Omega}\big{(}\alpha(|\nabla\phi|)-\alpha(|\nabla\tilde{\phi}|)\big{)}\nabla\phi\cdot\nabla\tau\,{\rm d}\Omega\qquad\forall~{}~{}\tau\in H_{0}^{1}(\Omega).

Taking τ=ϕ~ϕ\tau=\tilde{\phi}-\phi in the above equation and using Lemma 2.2 give

(ϕ~ϕ)2\displaystyle\|\nabla(\tilde{\phi}-\phi)\|^{2} a(ϕ~;ϕ~ϕ,ϕ~ϕ)\displaystyle\leq a(\tilde{\phi};\tilde{\phi}-\phi,\tilde{\phi}-\phi)
=a(ϕ;ϕ,ϕ~ϕ)a(ϕ~;ϕ,ϕ~ϕ)\displaystyle=a(\phi;\phi,\tilde{\phi}-\phi)-a(\tilde{\phi};\phi,\tilde{\phi}-\phi)
CαϕL(ϕ~ϕ)2,\displaystyle\leq C_{\alpha}\|\nabla\phi\|_{L^{\infty}}\|\nabla(\tilde{\phi}-\phi)\|^{2},

which, together with (22), implies ϕ~=ϕ\tilde{\phi}=\phi. This finishes the proof. ∎

For equation (18), we have the following conclusion:

Theorem 2.6.

For any given ϕH01(Ω)\phi\in H_{0}^{1}(\Omega), equation (18) admits a unique solution 𝐇=ϕ𝐇0(curl)\boldsymbol{H}=\nabla\phi\in\boldsymbol{H}_{0}({\rm curl\,}), which means curl𝐇=0.{\rm curl\,}\boldsymbol{H}=0.

Proof.

The de Rham complex [6] on 2D2D and 3D3D domain implies that ϕ𝑯0(curl)\nabla\phi\in\boldsymbol{H}_{0}({\rm curl\,}). Thus, taking 𝑪=𝑯ϕ𝑯0(curl)\boldsymbol{C}=\boldsymbol{H}-\nabla\phi\in\boldsymbol{H}_{0}({\rm curl\,}) in (18) yields 𝑯=ϕ.\boldsymbol{H}=\nabla\phi.

Since (19)-(20) are the weak formulations of the Navier-Stokes equations, the following existence and uniqueness results are standard (cf. [13, Page 285-287, Theorems 2.1 and 2.2]).

Theorem 2.7.

Given 𝐟(H1(Ω))d\boldsymbol{f}\in(H^{-1}(\Omega))^{d}, there exists at least one pair (𝐮,p~)𝐇01(Ω)×L02(Ω)(\boldsymbol{u},\tilde{p})\in\boldsymbol{H}_{0}^{1}(\Omega)\times L_{0}^{2}(\Omega) satisfying (19) - (20), and holds

𝒖1η𝒇1,\|\nabla\boldsymbol{u}\|\leq\frac{1}{\eta}\|\boldsymbol{f}\|_{-1},

where 𝐟1:=sup𝐯𝐇01(Ω)Ω𝐟𝐯dΩ𝐯.\|\boldsymbol{f}\|_{-1}:=\sup\limits_{\boldsymbol{v}\in\boldsymbol{H}_{0}^{1}(\Omega)}\frac{\int_{\Omega}\boldsymbol{f}\cdot\boldsymbol{v}\,{\rm d}\Omega}{\|\nabla\boldsymbol{v}\|}. In addition, if

(ρ𝒩/η2)𝒇1<1with𝒩:=sup𝒖,𝒗,𝒘𝑯01(Ω)Ω(𝒖)𝒗𝒘dΩ𝒖𝒗𝒘,(\rho\mathcal{N}/\eta^{2})\|\boldsymbol{f}\|_{-1}<1\quad\text{with}\quad\mathcal{N}:=\sup\limits_{\boldsymbol{u},\boldsymbol{v},\boldsymbol{w}\in\boldsymbol{H}_{0}^{1}(\Omega)}\frac{\int_{\Omega}(\boldsymbol{u}\cdot\nabla)\boldsymbol{v}\cdot\boldsymbol{w}\,{\rm d}\Omega}{\|\nabla\boldsymbol{u}\|\|\nabla\boldsymbol{v}\|\|\nabla\boldsymbol{w}\|},

then the solution pair (𝐮,p~)(\boldsymbol{u},\tilde{p}) is unique.

3. Finite element methods

3.1. Finite element spaces

Let 𝒯h\mathcal{T}_{h} be a quasi-uniform simplicial decomposition of Ω\Omega with mesh size h:=maxK𝒯hhKh:=\max\limits_{K\in\mathcal{T}_{h}}h_{K}, where hKh_{K} denotes the diameter of KK for any K𝒯hK\in\mathcal{T}_{h}.

For an integer l0l\geq 0, let l(K)\mathbb{P}_{l}(K) denote the set of polynomials, defined on K𝒯hK\in\mathcal{T}_{h}, of degree no more than ll. We introduce the following finite element spaces:

  • 𝑽h=(Vh)d(or (H01(Ω))d)\boldsymbol{V}_{h}=(V_{h})^{d}\ \big{(}\subset\text{or }\not\subset(H_{0}^{1}(\Omega))^{d}\big{)} is the Lagrange element space [13, 9, 14] for the velocity 𝒖\boldsymbol{u}, with l+1(K)Vh|K\mathbb{P}_{l+1}(K)\subset V_{h}|_{K} for any K𝒯hK\in\mathcal{T}_{h}. In particular, for the nonconforming case that VhH01(Ω)V_{h}\not\subset H_{0}^{1}(\Omega), each vhVhv_{h}\in V_{h} is required to satisfy the following conditions:

    • (i)

      vhv_{h} vanishes at all the nodes on Ω\partial\Omega;

    • (ii)

      |vh|1,h:=(K𝒯hvhK2)1/2|v_{h}|_{1,h}:=(\sum\limits_{K\in\mathcal{T}_{h}}\|\nabla v_{h}\|_{K}^{2})^{1/2} is a norm.

    Note that the classical nonconforming Crouzeix-Raviart (CR) finite element  [12] is corresponding to the nonconforming case with l=0l=0. In the nonconforming case, the gradient and divergence operators, \nabla and \nabla\cdot, in the finite element scheme (28) - (29) will be understood as h\nabla_{h} and h\nabla_{h}\cdot, respectively, which denote respectively the piecewise gradient and divergence operators acting on element by element in 𝒯h\mathcal{T}_{h}.

  • WhL02(Ω)W_{h}\subset L_{0}^{2}(\Omega) is the piecewise polynomial space for the ‘pressure’ variable p~\tilde{p}, with l(K)Wh|K\mathbb{P}_{l}(K)\subset W_{h}|_{K} for any K𝒯hK\in\mathcal{T}_{h}.

  • ShH01(Ω)S_{h}\subset H_{0}^{1}(\Omega) is the Lagrange element space [11] for the new variable ϕ\phi, with l+1(K)Sh|K\mathbb{P}_{l+1}(K)\subset S_{h}|_{K} for any K𝒯hK\in\mathcal{T}_{h}.

  • 𝑼h𝑯0(curl)\boldsymbol{U}_{h}\subset\boldsymbol{H}_{0}({\rm curl\,}) is the edge element space [18, 19] for the magnetic field 𝑯\boldsymbol{H}, with (l(K))d𝑼h|K(\mathbb{P}_{l}(K))^{d}\subset\boldsymbol{U}_{h}|_{K} and (l(K))2d3curl𝑼h|K(\mathbb{P}_{l}(K))^{2d-3}\subset{\rm curl\,}\boldsymbol{U}_{h}|_{K} for any K𝒯hK\in\mathcal{T}_{h}.

In addition, we make the following assumptions for the above finite element spaces.

  • (A1)

    There holds the inf-sup condition

    (23) inf𝒗h𝑽h(𝒗h,qh)𝒗h1β0qhqhWh,\inf\limits_{\boldsymbol{v}_{h}\in\boldsymbol{V}_{h}}\frac{(\nabla\cdot\boldsymbol{v}_{h},q_{h})}{\|\boldsymbol{v}_{h}\|_{1}}\geq\beta_{0}\|q_{h}\|\qquad\forall~{}~{}q_{h}\in W_{h},

    where β0>0\beta_{0}>0 is a constant independent of hh;

  • (A2)

    The diagram

    (24) H01grad𝑯0(curl)πhs@ VVπhcVShgrad𝑼h\begin{CD}H_{0}^{1}@>{{\rm grad\,}}>{}>\boldsymbol{H}_{0}({\rm curl\,})\\ @V{}V{\pi_{h}^{s}}V@ VV\pi^{c}_{h}V\\ \ S_{h}@>{{\rm grad\,}}>{}>\boldsymbol{U}_{h}\end{CD}

    is a commutative sequence. Here πhs:H01(Ω)C0(Ω¯)Sh\pi_{h}^{s}:~{}H_{0}^{1}(\Omega)\cap C^{0}(\overline{\Omega})\rightarrow S_{h} and πhc:𝑯0(curl)𝑼h\pi_{h}^{c}:~{}\boldsymbol{H}_{0}({\rm curl\,})\rightarrow\boldsymbol{U}_{h} are the classical interpolation operators, and grad{\rm grad\,} denotes the gradient operator. Note that the diagram (24) also indicates that

    gradSh𝑼h.{\rm grad\,}S_{h}\subset\boldsymbol{U}_{h}.

We recall the following inverse inequality:

(25) shCinvhd/2shshSh,\|\nabla s_{h}\|_{\infty}\leq C_{inv}h^{-d/2}\|\nabla s_{h}\|\qquad\forall s_{h}\in S_{h},

where Cinv>0C_{inv}>0 is a constant independent of hh.

Remark 3.1.

There are many finite element spaces that satisfy (A1) and (A2). For example, we can choose 𝐕h\boldsymbol{V}_{h} and WhW_{h} as the Taylor-Hood element pairs. And the spaces ShS_{h} and 𝐔h\boldsymbol{U}_{h} can be respectively chosen as the lowest order Lagrange finite element space and the lowest order Nédélec edge element space  [5, 6, 7].

Remark 3.2.

In this paper, we consider the conforming finite element spaces for the Navier-Stokes equation for simply of notations. In fact the nonconforming finite element spaces which satisfying the inf-sup condition (A1), with replace the global differential operators to element wise, are also work for the Navier-Stokes equations. For example the CRCR - 0\mathbb{P}_{0} finite element spaces are work for the Navier-Stokes equations.

3.2. Finite element scheme

In view of (17)-(20), we consider the following finite element scheme for the FHD model: Find ϕhSh\phi_{h}\in S_{h}, 𝑯h𝑼h\boldsymbol{H}_{h}\in\boldsymbol{U}_{h}, 𝒖h𝑽h\boldsymbol{u}_{h}\in\boldsymbol{V}_{h} and p~hWh\tilde{p}_{h}\in W_{h}, such that

(26) a(ϕh;ϕh,τh)=(g,τh)\displaystyle a(\phi_{h};\phi_{h},\tau_{h})=-(g,\tau_{h}) τhSh,\displaystyle\forall~{}~{}\tau_{h}\in S_{h},
(27) (𝑯h,𝑪h)(ϕh,𝑪h)=0\displaystyle(\boldsymbol{H}_{h},\boldsymbol{C}_{h})-(\nabla\phi_{h},\boldsymbol{C}_{h})=0 𝑪h𝑼h,\displaystyle\forall~{}~{}\boldsymbol{C}_{h}\in\boldsymbol{U}_{h},
(28) η(𝒖h,𝒗h)+b(𝒖h;𝒖h,𝒗h)(p~h,𝒗h)=(𝒇,𝒗h)\displaystyle\eta(\nabla\boldsymbol{u}_{h},\nabla\boldsymbol{v}_{h})+b(\boldsymbol{u}_{h};\boldsymbol{u}_{h},\boldsymbol{v}_{h})-(\tilde{p}_{h},\nabla\cdot\boldsymbol{v}_{h})=(\boldsymbol{f},\boldsymbol{v}_{h}) 𝒗h𝑽h,\displaystyle\forall~{}~{}\boldsymbol{v}_{h}\in\boldsymbol{V}_{h},
(29) (𝒖h,qh)=0\displaystyle(\nabla\cdot\boldsymbol{u}_{h},q_{h})=0\ qhWh,\displaystyle\forall~{}~{}q_{h}\in W_{h},

where the trilinear form b(;,):𝑯1(Ω)×𝑯1(Ω)×𝑯1(Ω)b(\cdot;\cdot,\cdot):\boldsymbol{H}^{1}(\Omega)\times\boldsymbol{H}^{1}(\Omega)\times\boldsymbol{H}^{1}(\Omega)\rightarrow\mathbb{R} is defined by

b(𝒘;𝒖,𝒗):=ρ2[((𝒘)𝒖,𝒗)((𝒘)𝒗,𝒖)].b(\boldsymbol{w};\boldsymbol{u},\boldsymbol{v}):=\frac{\rho}{2}[((\boldsymbol{w}\cdot\nabla)\boldsymbol{u},\boldsymbol{v})-((\boldsymbol{w}\cdot\nabla)\boldsymbol{v},\boldsymbol{u})].
Remark 3.3.

Similar to (17)-(20), the finite element scheme (26) - (29) is a decoupled system. We can first solve the nonlinear equation (26) to get ϕh\phi_{h}, and solve the Navier-Stokes system (28) - (29) to get 𝐮h\boldsymbol{u}_{h} and p~h\tilde{p}_{h}. Then we can get 𝐇h\boldsymbol{H}_{h} from (27). Finally, we can get 𝐌h𝐔h\boldsymbol{M}_{h}\in\boldsymbol{U}_{h}, the approximation of the magnetization 𝐌\boldsymbol{M}, from

(30) (𝑴h,𝑭h)=((α(Hh)1)𝑯h,𝑭h)𝑭h𝑼h(\boldsymbol{M}_{h},\boldsymbol{F}_{h})=\big{(}(\alpha(H_{h})-1)\boldsymbol{H}_{h},\boldsymbol{F}_{h}\big{)}\,\,\,\,\,\qquad\quad\forall~{}~{}\boldsymbol{F}_{h}\in\boldsymbol{U}_{h}

with Hh:=|𝐇h|H_{h}:=|\boldsymbol{H}_{h}|, get ψhWh\psi_{h}\in W_{h}, the approximation of ψ\psi, from

(31) (ψh,χh)=(β(Hh),χh)χhWh,(\psi_{h},\chi_{h})=\big{(}\beta(H_{h}),\chi_{h}\big{)}\ \ \qquad\qquad\qquad\forall~{}~{}\chi_{h}\in W_{h},

and get phWhp_{h}\in W_{h}, the approximation of the pressure pp, from

(32) ph=p~h+μ0ψh.p_{h}=\tilde{p}_{h}+\mu_{0}\psi_{h}.
Remark 3.4.

It is easy to see that the trilinear form b(;,)b(\cdot;\cdot,\cdot) is skew-symmetric with respect to the last two variables, i.e.,

(33) b(𝒘;𝒖,𝒗)=b(𝒘;𝒗,𝒖)𝒘,𝒖,𝒗𝑯1(Ω),b(\boldsymbol{w};\boldsymbol{u},\boldsymbol{v})=-b(\boldsymbol{w};\boldsymbol{v},\boldsymbol{u})\qquad\forall~{}\boldsymbol{w},~{}\boldsymbol{u},~{}\boldsymbol{v}\in\boldsymbol{H}^{1}(\Omega),

and that

(34) b(𝒘;𝒖,𝒗)=ρ((𝒘)𝒖,𝒗)+ρ2((𝒘)𝒖,𝒗)𝒘,𝒖𝑯1(Ω),𝒗𝑯01(Ω).b(\boldsymbol{w};\boldsymbol{u},\boldsymbol{v})=\rho\big{(}(\boldsymbol{w}\cdot\nabla)\boldsymbol{u},\boldsymbol{v}\big{)}+\frac{\rho}{2}\big{(}(\nabla\cdot\boldsymbol{w})\boldsymbol{u},\boldsymbol{v}\big{)}\quad\forall~{}\boldsymbol{w},~{}\boldsymbol{u}\in\boldsymbol{H}^{1}(\Omega),\ \ \forall\ \boldsymbol{v}\in\boldsymbol{H}_{0}^{1}(\Omega).

As a result, the following two relations hold:

(35) b(𝒘;𝒗,𝒗)=0𝒘,𝒗𝑯1(Ω),b(\boldsymbol{w};\boldsymbol{v},\boldsymbol{v})=0\qquad\forall~{}\boldsymbol{w},\boldsymbol{v}\in\boldsymbol{H}^{1}(\Omega),\\
(36) b(𝒘;𝒖,𝒗)=ρ((𝒘)𝒖,𝒗)𝒘,𝒖𝑯1(Ω) with div𝒘=0,𝒗𝑯01(Ω).b(\boldsymbol{w};\boldsymbol{u},\boldsymbol{v})=\rho\big{(}(\boldsymbol{w}\cdot\nabla)\boldsymbol{u},\boldsymbol{v}\big{)}\quad\forall~{}\boldsymbol{w},~{}\boldsymbol{u}\in\boldsymbol{H}^{1}(\Omega)\text{ with }\operatorname{div}\boldsymbol{w}=0,\ \forall\ \boldsymbol{v}\in\boldsymbol{H}_{0}^{1}(\Omega).

Theorems 3.5-3.7 show the wellposedness of the finite element scheme (26) - (29).

Theorem 3.5.

The nonlinear discrete equation (26) has at least one solution ϕhSh\phi_{h}\in S_{h}, and there holds

(37) ϕh1μ0𝑯e.\|\nabla\phi_{h}\|\leq\frac{1}{\mu_{0}}\|\boldsymbol{H}_{e}\|.

Furthermore, if the external magnetic filed 𝐇e\boldsymbol{H}_{e} satisfies

(38) 𝑯e<μ0Cinv1Cα1hd/2,\|\boldsymbol{H}_{e}\|<\mu_{0}C_{inv}^{-1}C_{\alpha}^{-1}h^{d/2},

then (26) admits a unique solution ϕhSh\phi_{h}\in S_{h}, and there holds

(39) ϕh<1/Cα.\|\nabla\phi_{h}\|_{\infty}<1/C_{\alpha}.
Proof.

Define an operator A:ShShA:~{}S_{h}\rightarrow S_{h}^{\prime} by

A(ϕh),τh=a(ϕh;ϕh,τh)ϕh,τhSh,\langle A(\phi_{h}),\tau_{h}\rangle=a(\phi_{h};\phi_{h},\tau_{h})\qquad\forall~{}~{}\phi_{h},~{}\tau_{h}\in S_{h},

and define Φ:ShSh\Phi:~{}S_{h}\rightarrow S_{h}^{\prime} by

Φ(ϕh)=A(ϕh)+QhgϕhSh,\Phi(\phi_{h})=A(\phi_{h})+Q_{h}^{\prime}g\qquad\forall~{}~{}\phi_{h}\in S_{h},

where Qh:(H01(Ω))ShQ_{h}^{\prime}:~{}(H_{0}^{1}(\Omega))^{\prime}\rightarrow S_{h}^{\prime} is given by

(40) Qhg,τh=g,τh=1μ0(𝑯e,τh)τhSh.\langle Q_{h}^{\prime}g,\tau_{h}\rangle=\langle g,\tau_{h}\rangle=-\frac{1}{\mu_{0}}(\boldsymbol{H}_{e},\nabla\tau_{h})\qquad\forall~{}~{}\tau_{h}\in S_{h}.

Thus, (26) is equivalent to the equation

Φ(ϕh)=0.\Phi(\phi_{h})=0.

Lemma 2.4 implies

A(ϕh),ϕh=a(ϕh;ϕh,ϕh)ϕh2ϕhSh.\langle A(\phi_{h}),\phi_{h}\rangle=a(\phi_{h};\phi_{h},\phi_{h})\geq\|\nabla\phi_{h}\|^{2}\qquad\forall~{}~{}\phi_{h}\in S_{h}.

The definition (40) and the Cauchy-Schwarz inequality imply

Qhg,ϕh=g,ϕh1μ0𝑯eϕhϕhSh.\langle Q_{h}^{\prime}g,\phi_{h}\rangle=\langle g,\phi_{h}\rangle\leq\frac{1}{\mu_{0}}\|\boldsymbol{H}_{e}\|\|\nabla\phi_{h}\|\qquad\forall~{}~{}\phi_{h}\in S_{h}.

Therefore, we have

Φ(ϕh),ϕh=A(ϕh)+Qhg,ϕh0,ϕhSh with ϕh=1μ0𝑯e.\langle\Phi(\phi_{h}),\phi_{h}\rangle=\langle A(\phi_{h})+Q_{h}^{\prime}g,\phi_{h}\rangle\geq 0,\qquad\forall~{}~{}\phi_{h}\in S_{h}\text{ with }\|\nabla\phi_{h}\|=\frac{1}{\mu_{0}}\|\boldsymbol{H}_{e}\|.

For any given 0ϕh0Sh0\neq\phi_{h0}\in S_{h} and ϵ>0\epsilon>0, denote δ:=min{ϵ2C1,ϵ2Cαϕh0}\delta:=\min\left\{\frac{\epsilon}{2C_{1}},~{}\frac{\epsilon}{2C_{\alpha}\|\nabla\phi_{h0}\|_{\infty}}\right\}. Then for any ϕhSh\phi_{h}\in S_{h} satisfying (ϕhϕh0)<δ\|\nabla(\phi_{h}-\phi_{h0})\|<\delta, by Lemmas 2.2 and 2.4 we have

Φ(ϕh)Φ(ϕh0)Sh=A(ϕh)A(ϕh0)Sh\displaystyle\|\Phi(\phi_{h})-\Phi(\phi_{h0})\|_{S_{h}^{\prime}}=\|A(\phi_{h})-A(\phi_{h0})\|_{S_{h}^{\prime}}
=\displaystyle= supτhShA(ϕh)A(ϕh0),τhτh\displaystyle\sup\limits_{\tau_{h}\in S_{h}}\frac{\langle A(\phi_{h})-A(\phi_{h0}),\tau_{h}\rangle}{\|\nabla\tau_{h}\|}
=\displaystyle= supτhSha(ϕh;ϕh,τh)a(ϕh0;ϕh0,τh)τh\displaystyle\sup\limits_{\tau_{h}\in S_{h}}\frac{a(\phi_{h};\phi_{h},\tau_{h})-a(\phi_{h0};\phi_{h0},\tau_{h})}{\|\nabla\tau_{h}\|}
=\displaystyle= supτhSha(ϕh;ϕh,τh)a(ϕh;ϕh0,τh)+a(ϕh;ϕh0,τh)a(ϕh0;ϕh0,τh)τh\displaystyle\sup\limits_{\tau_{h}\in S_{h}}\frac{a(\phi_{h};\phi_{h},\tau_{h})-a(\phi_{h};\phi_{h0},\tau_{h})+a(\phi_{h};\phi_{h0},\tau_{h})-a(\phi_{h0};\phi_{h0},\tau_{h})}{\|\nabla\tau_{h}\|}
\displaystyle\leq C1(ϕhϕh0)+Cαϕh0(ϕhϕh0)\displaystyle C_{1}\|\nabla(\phi_{h}-\phi_{h0})\|+C_{\alpha}\|\nabla\phi_{h0}\|_{\infty}\|\nabla(\phi_{h}-\phi_{h0})\|
<\displaystyle< ϵ.\displaystyle\epsilon.

This means that Φ\Phi is continuous on the set Sh{0}S_{h}\setminus\{0\}. Notice that Lemma 2.4 implies Φ\Phi is continuous at the point 0. Hence, Φ\Phi is continuous on ShS_{h}.

On the other hand, the Riesz representation theorem implies that the spaces ShS_{h} and ShS_{h}^{\prime} are isometry. As a result, by [13, Page 279, Corollary 1.1] equation (26) admits at least one solution ϕhSh\phi_{h}\in S_{h} satisfiying (37).

If 𝑯e\boldsymbol{H}_{e} satisfies assumption (38), then from (37) and the inverse inequality (25) we get

ϕhCinvhd/2ϕh<1/Cα,\|\nabla\phi_{h}\|_{\infty}\leq C_{inv}h^{-d/2}\|\nabla\phi_{h}\|<1/C_{\alpha},

i.e. (39) holds. Thus, by following the same line as in the proof of the uniqueness of the weak solution ϕ\phi in Theorem 2.5, we can easily obtain the uniqueness of the discrete solution ϕh\phi_{h}. ∎

Theorem 3.6.

For any given ϕhSh\phi_{h}\in S_{h}, equation (27) admits a unique solution 𝐇h=ϕh𝐔h\boldsymbol{H}_{h}=\nabla\phi_{h}\in\boldsymbol{U}_{h}, which means curl𝐇h=0.{\rm curl\,}\boldsymbol{H}_{h}=0.

Proof.

For any given ϕhSh\phi_{h}\in S_{h}, assumption (A2) on the finite element spaces implies that ϕh𝑼h\nabla\phi_{h}\in\boldsymbol{U}_{h}. Thus, taking 𝑪h=𝑯hϕh𝑼h\boldsymbol{C}_{h}=\boldsymbol{H}_{h}-\nabla\phi_{h}\in\boldsymbol{U}_{h} in (27) yields 𝑯h=ϕh.\boldsymbol{H}_{h}=\nabla\phi_{h}.

The following well-posedness results for the finite element scheme (28) - (29) of the Navier-Stokes equations are standard (cf. [13]).

Theorem 3.7.

The finite element scheme (28) - (29) has at least one solution (𝐮h,p~h)𝐕h×Wh(\boldsymbol{u}_{h},\tilde{p}_{h})\in\boldsymbol{V}_{h}\times W_{h}, and there holds

𝒖h1η𝒇1.\|\nabla\boldsymbol{u}_{h}\|\leq\frac{1}{\eta}\|\boldsymbol{f}\|_{-1}.

Further more, if

(41) (𝒩~/η2)𝒇1<1with𝒩~=sup𝒘h,𝒖h,𝒗h𝑽hb(𝒘h;𝒖h,𝒗h)𝒘h𝒖h𝒗h,(\tilde{\mathcal{N}}/\eta^{2})\|\boldsymbol{f}\|_{-1}<1\quad\text{with}\quad\tilde{\mathcal{N}}=\sup\limits_{\boldsymbol{w}_{h},\boldsymbol{u}_{h},\boldsymbol{v}_{h}\in\boldsymbol{V}_{h}}\frac{b(\boldsymbol{w}_{h};\boldsymbol{u}_{h},\boldsymbol{v}_{h})}{\|\nabla\boldsymbol{w}_{h}\|\|\nabla\boldsymbol{u}_{h}\|\|\nabla\boldsymbol{v}_{h}\|},

then the finite element solution (𝐮h,p~h)(\boldsymbol{u}_{h},\tilde{p}_{h}) is unique.

3.3. Error analysis

In this subsection, we will give some error estimates for the finite element scheme (26) - (29), under some rational assumptions.

Firstly, we have the following error estimate for the discrete solution ϕh\phi_{h}.

Theorem 3.8.

Let ϕH01(Ω)Hl+2(Ω)\phi\in H_{0}^{1}(\Omega)\cap H^{l+2}(\Omega) be the solution of (17), and let ϕhSh\phi_{h}\in S_{h} be the solution of (26). Assume that ϕ\phi satisfies (22), i.e. ϕL<1/Cα,\|\nabla\phi\|_{L^{\infty}}<1/C_{\alpha}, then there exists a positive constant CsC_{s}, independent of hh, such that

(ϕϕh)Cs(1+C1)1CαϕLhl+1ϕl+2.\|\nabla(\phi-\phi_{h})\|\leq\frac{C_{s}(1+C_{1})}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}h^{l+1}\|\phi\|_{l+2}.
Proof.

Let ϕhSh\phi_{h}^{*}\in S_{h} be the classical interpolation of ϕ\phi, satisfying the estimate

(42) (ϕϕh)Cshl+1ϕl+2,\|\nabla(\phi-\phi_{h}^{*})\|\leq C_{s}h^{l+1}\|\phi\|_{l+2},

where Cs>0C_{s}>0 is a constant independent of hh and ϕ\phi.

In view of the inequality

(ϕϕh)(ϕϕh)+(ϕhϕh),\|\nabla(\phi-\phi_{h})\|\leq\|\nabla(\phi-\phi_{h}^{*})\|+\|\nabla(\phi_{h}^{*}-\phi_{h})\|,

it suffices to estimate the term (ϕhϕh)\|\nabla(\phi_{h}^{*}-\phi_{h})\|. To this end, we apply Lemma 2.4 and the relation

a(ϕ;ϕ,τh)=a(ϕh;ϕh,τh),τhSh,a(\phi;\phi,\tau_{h})=a(\phi_{h};\phi_{h},\tau_{h}),\quad\forall\ \tau_{h}\in S_{h},

to get

(ϕhϕh)2\displaystyle\|\nabla(\phi_{h}^{*}-\phi_{h})\|^{2} \displaystyle\leq a(ϕh;ϕhϕh,ϕhϕh)\displaystyle a(\phi_{h};\phi_{h}^{*}-\phi_{h},\phi_{h}^{*}-\phi_{h})
=\displaystyle= a(ϕh;ϕhϕ,ϕhϕh)+a(ϕh;ϕϕh,ϕhϕh).\displaystyle a(\phi_{h};\phi_{h}^{*}-\phi,\phi_{h}^{*}-\phi_{h})+a(\phi_{h};\phi-\phi_{h},\phi_{h}^{*}-\phi_{h}).

Since

|a(ϕh;ϕhϕ,ϕhϕh)|C1(ϕϕh)(ϕhϕh)|a(\phi_{h};\phi_{h}^{*}-\phi,\phi_{h}^{*}-\phi_{h})|\leq C_{1}\|\nabla(\phi-\phi_{h}^{*})\|\|\nabla(\phi_{h}^{*}-\phi_{h})\|

and

a(ϕh;ϕϕh,ϕhϕh)=\displaystyle a(\phi_{h};\phi-\phi_{h},\phi_{h}^{*}-\phi_{h})= a(ϕh;ϕ,ϕhϕh)a(ϕ;ϕ,ϕhϕh)\displaystyle a(\phi_{h};\phi,\phi_{h}^{*}-\phi_{h})-a(\phi;\phi,\phi_{h}^{*}-\phi_{h})
=\displaystyle= ((α(|ϕh|)α(|ϕ|))ϕ,(ϕhϕh))\displaystyle\left(\big{(}\alpha(|\nabla\phi_{h}|)-\alpha(|\nabla\phi|)\big{)}\nabla\phi,\nabla(\phi_{h}^{*}-\phi_{h})\right)
\displaystyle\leq CαϕL(ϕϕh)(ϕhϕh)\displaystyle C_{\alpha}\|\nabla\phi\|_{L^{\infty}}\|\nabla(\phi-\phi_{h})\|\|\nabla(\phi_{h}^{*}-\phi_{h})\|
\displaystyle\leq CαϕL((ϕϕh)+(ϕhϕh))(ϕhϕh),\displaystyle C_{\alpha}\|\nabla\phi\|_{L^{\infty}}\big{(}\|\nabla(\phi-\phi_{h}^{*})\|+\|\nabla(\phi_{h}^{*}-\phi_{h})\|\big{)}\|\nabla(\phi_{h}^{*}-\phi_{h})\|,

we have

(ϕhϕh)C1(ϕϕh)+CαϕL((ϕϕh)+(ϕhϕh)),\displaystyle\|\nabla(\phi_{h}^{*}-\phi_{h})\|\leq C_{1}\|\nabla(\phi-\phi_{h}^{*})\|+C_{\alpha}\|\nabla\phi\|_{L^{\infty}}\big{(}\|\nabla(\phi-\phi_{h}^{*})\|+\|\nabla(\phi_{h}^{*}-\phi_{h})\|\big{)},

which, together with (22), implies

(ϕhϕh)C1+CαϕL1CαϕL(ϕϕh).\displaystyle\|\nabla(\phi_{h}^{*}-\phi_{h})\|\leq\frac{C_{1}+C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}\|\nabla(\phi-\phi_{h}^{*})\|.

This inequality, together with (42), indicates the desired conclusion. ∎

In light of Theorems 2.6 and 3.6, we know that

𝑯=ϕ,𝑯h=ϕhandcurl𝑯=curl𝑯h=0.\boldsymbol{H}=\nabla\phi,\quad\boldsymbol{H}_{h}=\nabla\phi_{h}\quad\text{and}\quad{\rm curl\,}\boldsymbol{H}={\rm curl\,}\boldsymbol{H}_{h}=0.

Thus, by Theorem 3.8, we immediately get the following error estimate for the discrete solution 𝑯h\boldsymbol{H}_{h} of (27).

Theorem 3.9.

Let 𝐇𝐇0(curl)\boldsymbol{H}\in\boldsymbol{H}_{0}({\rm curl\,}) be the solution of (18), and let 𝐇h𝐔h\boldsymbol{H}_{h}\in\boldsymbol{U}_{h} be the solution of (27). Then, under the same conditions as in Theorem 3.8, there holds

𝑯𝑯h+curl(𝑯𝑯h)Cs(1+C1)1CαϕLhl+1ϕl+2.\|\boldsymbol{H}-\boldsymbol{H}_{h}\|+\|{\rm curl\,}(\boldsymbol{H}-\boldsymbol{H}_{h})\|\leq\frac{C_{s}(1+C_{1})}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}h^{l+1}\|\phi\|_{l+2}.

The following error estimate for the Navier-Stokes system (28)-(29) is standard (cf. [13]).

Theorem 3.10.

Let 𝐮𝐇01(Ω)𝐇l+2(Ω)\boldsymbol{u}\in\boldsymbol{H}_{0}^{1}(\Omega)\cap\boldsymbol{H}^{l+2}(\Omega) and p~L02(Ω)Hl+1(Ω)\tilde{p}\in L_{0}^{2}(\Omega)\cap H^{l+1}(\Omega) be the solution of the Navier-Stokes system (19)-(20), and let 𝐮h𝐕h\boldsymbol{u}_{h}\in\boldsymbol{V}_{h} and p~hWh\tilde{p}_{h}\in W_{h} be the solution of (28)-(29). There exists a positive constant CC, independent of hh, such that

𝒖𝒖h1+p~p~hChl+1(𝒖l+2+p~l+1).\|\boldsymbol{u}-\boldsymbol{u}_{h}\|_{1}+\|\tilde{p}-\tilde{p}_{h}\|\leq Ch^{l+1}(\|\boldsymbol{u}\|_{l+2}+\|\tilde{p}\|_{l+1}).

As shown in Remark 3.3, for the the magnetization 𝑴=(α(H)1)𝑯\boldsymbol{M}=(\alpha(H)-1)\boldsymbol{H}, the new variable ψ=β(H)\psi=\beta(H), and the pressure p=p~+μ0ψp=\tilde{p}+\mu_{0}\psi, we can obtain their approximations 𝑴h,ψh\boldsymbol{M}_{h},\psi_{h} and php_{h} from (30), (31) and (32), respectively. In view of Theorems 3.9 and 3.10, we easily get the following error estimates for these three approximation solutions.

Corollary 3.11.

Assume that 𝐌𝐇l+1(Ω)\boldsymbol{M}\in\boldsymbol{H}^{l+1}(\Omega) and ψHl+1(Ω)\psi\in H^{l+1}(\Omega). Under the same conditions as in Theorems 3.8 and 3.10, there exists positive constants CMC_{M} and CψC_{\psi}, independent of hh, such that

(43) 𝑴𝑴hCMhl+1(𝑴l+1+(1+C1)(1+C1+Cαϕ)1CαϕLϕl+2),\displaystyle\|\boldsymbol{M}-\boldsymbol{M}_{h}\|\leq C_{M}h^{l+1}\left(\|\boldsymbol{M}\|_{l+1}+\frac{(1+C_{1})(1+C_{1}+C_{\alpha}\|\nabla\phi\|_{\infty})}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}\|\phi\|_{l+2}\right),
(44) ψψh(Cψψl+1+MsCs(1+C1)1CαϕLϕl+2)hl+1,\displaystyle\|\psi-\psi_{h}\|\leq\left(C_{\psi}\|\psi\|_{l+1}+\frac{M_{s}C_{s}(1+C_{1})}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}\|\phi\|_{l+2}\right)h^{l+1},
(45) pph(C𝒖l+2+Cp~l+1+μ0Cψψl+1+μ0MsCs(1+C1)1CαϕLϕl+2)hl+1.\displaystyle\|p-p_{h}\|\leq\left(C\|\boldsymbol{u}\|_{l+2}+C\|\tilde{p}\|_{l+1}+\mu_{0}C_{\psi}\|\psi\|_{l+1}+\frac{\mu_{0}M_{s}C_{s}(1+C_{1})}{1-C_{\alpha}\|\nabla\phi\|_{L^{\infty}}}\|\phi\|_{l+2}\right)h^{l+1}.
Proof.

Equation (30) implies that

𝑴h=Qhc((α(Hh)1)𝑯h),\boldsymbol{M}_{h}=Q_{h}^{c}((\alpha(H_{h})-1)\boldsymbol{H}_{h}),

where Qhc:𝑳2(Ω)𝑼hQ_{h}^{c}:~{}\boldsymbol{L}^{2}(\Omega)\rightarrow\boldsymbol{U}_{h} is the L2L^{2} orthogonal projection operator. Using the Langevin law (4), we have

𝑴𝑴h=𝑴Qhc𝑴+Qhc((α(H)1)𝑯(α(Hh)1)𝑯h).\displaystyle\boldsymbol{M}-\boldsymbol{M}_{h}=\boldsymbol{M}-Q_{h}^{c}\boldsymbol{M}+Q_{h}^{c}\big{(}(\alpha(H)-1)\boldsymbol{H}-(\alpha(H_{h})-1)\boldsymbol{H}_{h}\big{)}.

Thus, by Lemma 2.2 (i), the boundedness of projection QhcQ_{h}^{c}, and the relation 𝑯=ϕ\boldsymbol{H}=\nabla\phi, we get

𝑴𝑴h\displaystyle\|\boldsymbol{M}-\boldsymbol{M}_{h}\| 𝑴Qhc𝑴+(α(H)α(Hh))𝑯+(α(Hh)1)(𝑯𝑯h)\displaystyle\leq\|\boldsymbol{M}-Q_{h}^{c}\boldsymbol{M}\|+\|(\alpha(H)-\alpha(H_{h}))\boldsymbol{H}\|+\|(\alpha(H_{h})-1)(\boldsymbol{H}-\boldsymbol{H}_{h})\|
𝑴Qhc𝑴+(Cα𝑯+C1+1)𝑯𝑯h\displaystyle\leq\|\boldsymbol{M}-Q_{h}^{c}\boldsymbol{M}\|+(C_{\alpha}\|\boldsymbol{H}\|_{\infty}+C_{1}+1)\|\boldsymbol{H}-\boldsymbol{H}_{h}\|
𝑴Qhc𝑴+(Cαϕ+C1+1)𝑯𝑯h,\displaystyle\leq\|\boldsymbol{M}-Q_{h}^{c}\boldsymbol{M}\|+(C_{\alpha}\|\nabla\phi\|_{\infty}+C_{1}+1)\|\boldsymbol{H}-\boldsymbol{H}_{h}\|,

which, together with Theorem 3.9 and the standard error estimation of the projection, gives the desired estimate for 𝑴h\boldsymbol{M}_{h}.

Similarly, equation (31) implies that

𝝍h=Qhw(β(Hh)),\boldsymbol{\psi}_{h}=Q_{h}^{w}(\beta(H_{h})),

where Qhw:L2(Ω)WhQ_{h}^{w}:~{}L^{2}(\Omega)\rightarrow W_{h} is the L2L^{2} orthogonal projection. Then by Lemma 2.2 (ii) we obtain

ψψh\displaystyle\|\psi-\psi_{h}\| ψQhwψ+Qhw(β(H)β(HH))\displaystyle\leq\|\psi-Q_{h}^{w}\psi\|+\|Q_{h}^{w}\big{(}\beta(H)-\beta(H_{H})\big{)}\|
ψQhwψ+(β(H)β(Hh))\displaystyle\leq\|\psi-Q_{h}^{w}\psi\|+\|\big{(}\beta(H)-\beta(H_{h})\big{)}\|
ψQhwψ+MsHHh\displaystyle\leq\|\psi-Q_{h}^{w}\psi\|+M_{s}\|H-H_{h}\|
ψQhwψ+Ms𝑯𝑯h,\displaystyle\leq\|\psi-Q_{h}^{w}\psi\|+M_{s}\|\boldsymbol{H}-\boldsymbol{H}_{h}\|,

which, together with Theorem 3.9 and the projection property, yields the desired estimate (44).

Finally, (32) means that

pphp~p~h+μ0ψψh,\|p-p_{h}\|\leq\|\tilde{p}-\tilde{p}_{h}\|+\mu_{0}\|\psi-\psi_{h}\|,

which, together with Theorem 3.10 and estimate (44), indicates the desired result (45). This completes the proof. ∎

4. Numerical experiments

This section is devoted to three numerical examples to verify the performance of the mixed finite element methods. In all the examples, we solve the nonlinear system (26) - (32) by using the iFEM package [10] and Algorithm 4.1.

Algorithm 4.1.

Given ϕh0Sh\phi_{h}^{0}\in S_{h} and 𝐮h0𝐕h\boldsymbol{u}_{h}^{0}\in\boldsymbol{V}_{h}, find ϕhSh\phi_{h}\in S_{h}, 𝐇h𝐔h\boldsymbol{H}_{h}\in\boldsymbol{U}_{h}, 𝐌h𝐔h\boldsymbol{M}_{h}\in\boldsymbol{U}_{h}, 𝐮h𝐕h\boldsymbol{u}_{h}\in\boldsymbol{V}_{h}, p~hWh\tilde{p}_{h}\in W_{h}, ψhWh\psi_{h}\in W_{h}, and phWhp_{h}\in W_{h} through five steps:

  1. Step 1.

    For n=1,2,,Ln=1,2,\dots,L do

    1. (1)

      Solving the nonlinear system (26) as: Find ϕhnSh\phi_{h}^{n}\in S_{h} such that

      a(ϕhn1;ϕhn,τh)=(g,τh)τhSh.a(\phi_{h}^{n-1};\phi_{h}^{n},\tau_{h})=-(g,\tau_{h})\qquad\forall~{}\tau_{h}\in S_{h}.
    2. (2)

      Solving the nonlinear saddle point system (28)-(29) as: Find 𝒖hn𝑽h\boldsymbol{u}_{h}^{n}\in\boldsymbol{V}_{h} and p~hnWh\tilde{p}_{h}^{n}\in W_{h} such that

      η(𝒖hn,𝒗h)+b(𝒖hn1;𝒖hn,𝒗h)(p~hn,𝒗h)=(𝒇,𝒗h)\displaystyle\eta(\nabla\boldsymbol{u}_{h}^{n},\nabla\boldsymbol{v}_{h})+b(\boldsymbol{u}_{h}^{n-1};\boldsymbol{u}_{h}^{n},\boldsymbol{v}_{h})-(\tilde{p}_{h}^{n},\nabla\cdot\boldsymbol{v}_{h})=(\boldsymbol{f},\boldsymbol{v}_{h}) 𝒗h𝑽h,\displaystyle\forall~{}\boldsymbol{v}_{h}\in\boldsymbol{V}_{h},
      (𝒖hn,qh)=0\displaystyle(\nabla\cdot\boldsymbol{u}_{h}^{n},q_{h})=0 qhWh.\displaystyle\forall~{}q_{h}\in W_{h}.
  2. Step 2.

    Let ϕh=ϕhL\phi_{h}=\phi_{h}^{L}, 𝒖h=𝒖hL\boldsymbol{u}_{h}=\boldsymbol{u}_{h}^{L} and p~h=p~hL\tilde{p}_{h}=\tilde{p}_{h}^{L}.

  3. Step 3.

    Solving equation (27) to get 𝑯h𝑼h\boldsymbol{H}_{h}\in\boldsymbol{U}_{h}.

  4. Step 4.

    Solving equations (30) and (31) to get 𝑴h𝑼h\boldsymbol{M}_{h}\in\boldsymbol{U}_{h} and ψhWh\psi_{h}\in W_{h}, respectively.

  5. Step 5.

    Substituting p~h\tilde{p}_{h} and ψh\psi_{h} into (32) to get phWhp_{h}\in W_{h}.

Remark 4.2.

From the convergence theory of Newton-type methods [15, 28], we know that the iterative solution of Algorithm 4.1 will converge to the exact solution, provided that the iteration number LL is big enough and the initial guess is nearby the exact solution. In fact, in all the subsequent numerical examples we choose the initial guess ϕh0\phi_{h}^{0} as the corresponding finite element solution of the Poisson equation

Δϕ~=gin Ω\Delta\tilde{\phi}=g\qquad\text{in }\ \Omega

with the same boundary condition as that of the exact solution ϕ\phi, and choose the initial guess 𝐮h0\boldsymbol{u}_{h}^{0} as the corresponding finite element solution of the Stokes equations

{ηΔ𝒖~+p=𝒇in Ω,𝒖~=0in Ω\left\{\begin{array}[]{ll}-\eta\Delta\tilde{\boldsymbol{u}}+\nabla p=\boldsymbol{f}&\text{in }\Omega,\\ \nabla\cdot\tilde{\boldsymbol{u}}=0&\text{in }\Omega\end{array}\right.

with the same boundary condition as that of the exact solution 𝐮\boldsymbol{u}. In so doing, the choice L=2L=2 works well in the algorithm.

Refer to caption
Figure 1. The domain Ω=(0,1)2\Omega=(0,1)^{2}: 4×44\times 4 (left) and 8×88\times 8 (right).
Example 4.3.

This is a 2D test. We take Ω=(0,1)2\Omega=(0,1)^{2}, and use N×NN\times N uniform triangular mesh (cf. Figure 1) with N=4,8,16,32,64,128N=4,~{}8,~{}16,~{}32,~{}64,~{}128. The exact solution of the FHD model (16) is given by

ϕ(x,y)=(x2x)(y2y),\displaystyle\phi(x,y)=(x^{2}-x)(y^{2}-y),
𝑯(x,y)=((2x1)(y2y),(x2x)(2y1)),\displaystyle\boldsymbol{H}(x,y)=((2x-1)(y^{2}-y),(x^{2}-x)(2y-1))^{\intercal},
𝑴(x,y)=Ms(coth(γH)1γH)𝑯H,\displaystyle\boldsymbol{M}(x,y)=M_{s}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{\boldsymbol{H}}{H},
𝒖(x,y)=(sin(πy),sin(πx)),\displaystyle\boldsymbol{u}(x,y)=(\sin(\pi y),\sin(\pi x))^{\intercal},
p(x,y)=60x2y20y35,\displaystyle p(x,y)=60x^{2}y-20y^{3}-5,

with the parameters MsM_{s}, ρ\rho, η\eta, μ0\mu_{0} and γ\gamma all being chosen as 11.

We use the conforming linear (1\mathbb{P}_{1}) element to discretize the variable ϕ\phi, the lowest order edge element NE0NE_{0} to discretize the variables 𝐇\boldsymbol{H} and 𝐌\boldsymbol{M}, the CRCR (nonconforming-1\mathbb{P}_{1}) element to discretize the variables 𝐮\boldsymbol{u}, and the piecewise constant (0\mathbb{P}_{0}) element to discretize the variables p~\tilde{p}, ψ\psi and pp. Note that such a combination of finite element spaces corresponds to l=0l=0, then we easily see from Theorems 3.8 - 3.10 and Corollary 3.11 that the theoretical accuracy of the scheme is 𝒪(h)\mathcal{O}(h). Numerical results are listed in Table 1.

Table 1. Numerical results for Example 4.3.
NN (ϕϕh)ϕ\frac{\|\nabla(\phi-\phi_{h})\|}{\|\nabla\phi\|} 𝑯𝑯hc𝑯c\frac{\|\boldsymbol{H}-\boldsymbol{H}_{h}\|_{c}}{\|\boldsymbol{H}\|_{c}} 𝑴𝑴h𝑴\frac{\|\boldsymbol{M}-\boldsymbol{M}_{h}\|}{\|\boldsymbol{M}\|} 𝒖𝒖h1,h𝒖1\frac{\|\boldsymbol{u}-\boldsymbol{u}_{h}\|_{1,h}}{\|\boldsymbol{u}\|_{1}} pphp\frac{\|p-p_{h}\|}{\|p\|} curl𝑯h\|{\rm curl\,}\boldsymbol{H}_{h}\|_{\infty}
44 0.39430.3943 0.39430.3943 0.39410.3941 0.74240.7424 0.30830.3083 0.0003e120.0003e-12
88 0.20230.2023 0.20230.2023 0.20230.2023 0.43850.4385 0.15240.1524 0.0013e120.0013e-12
1616 0.10180.1018 0.10180.1018 0.10180.1018 0.23520.2352 0.07170.0717 0.0083e120.0083e-12
3232 0.05100.0510 0.05100.0510 0.05100.0510 0.12070.1207 0.03420.0342 0.0315e120.0315e-12
6464 0.02550.0255 0.02550.0255 0.02550.0255 0.06090.0609 0.01670.0167 0.1485e120.1485e-12
128128 0.01280.0128 0.01280.0128 0.01280.0128 0.03050.0305 0.00830.0083 0.6481e120.6481e-12
order 0.99000.9900 0.99000.9900 0.98990.9899 0.92070.9207 1.04391.0439
Refer to caption
Figure 2. The domain Ω=(0,1)3\Omega=(0,1)^{3}: 2×2×22\times 2\times 2 (left) and 4×4×44\times 4\times 4 (right).

The other two experiments, Examples 4.4 and 4.5, are 3D tests, with Ω=(0,1)3\Omega=(0,1)^{3} and use N×N×NN\times N\times N uniform tetrahedral meshes (cf. Figure 2) with N=4,8,16N=4,~{}8,~{}16.

Example 4.4 (The lowest order approximation).

In this 3D test, the exact solution of the FHD model (16) is given by

ϕ=sin(πx)sin(πy)sin(πz),\displaystyle\phi=\sin(\pi x)\sin(\pi y)\sin(\pi z),
𝑯=π(cos(πx)sin(πy)sin(πz)sin(πx)cos(πy)sin(πz)sin(πx)sin(πy)cos(πz)),\displaystyle\boldsymbol{H}=\pi\begin{pmatrix}\cos(\pi x)\sin(\pi y)\sin(\pi z)\\ \sin(\pi x)\cos(\pi y)\sin(\pi z)\\ \sin(\pi x)\sin(\pi y)\cos(\pi z)\end{pmatrix},
𝑴=Ms(coth(γH)1γH)𝑯H,\displaystyle\boldsymbol{M}=M_{s}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{\boldsymbol{H}}{H},
𝒖=(sin(πy),sin(πz),sin(πx)),\displaystyle\boldsymbol{u}=\begin{pmatrix}\sin(\pi y),&\sin(\pi z),&\sin(\pi x)\end{pmatrix}^{\intercal},
p=120x2yz40y3z40yz3,\displaystyle p=120x^{2}yz-40y^{3}z-40yz^{3},

with the parameters MsM_{s}, γ\gamma, ρ\rho and η\eta all being choosen as 11 and μ0=10\mu_{0}=10.

We use the the conforming 1\mathbb{P}_{1}-element to discretize the variable ϕ\phi, the lowest order Nédélec finite element space NE0NE_{0} [18, 19] to discretize the variables 𝐇\boldsymbol{H} and 𝐌\boldsymbol{M}, the nonconforming CRCR element [12] to discretize the variable 𝐮\boldsymbol{u}, and the piecewise constant 0\mathbb{P}_{0} element to discretize the variables p~\tilde{p}, ψ\psi and pp. With these settings, from Theorems 3.8 - 3.10 and Corollary 3.11 we see that the theoretical accuracy of the scheme is 𝒪(h)\mathcal{O}(h). Numerical results are given in Table 2.

Table 2. Numerical results for Example 4.4.
NN (ϕϕh)ϕ\frac{\|\nabla(\phi-\phi_{h})\|}{\|\nabla\phi\|} 𝑯𝑯hc𝑯c\frac{\|\boldsymbol{H}-\boldsymbol{H}_{h}\|_{c}}{\|\boldsymbol{H}\|_{c}} 𝑴𝑴h𝑴\frac{\|\boldsymbol{M}-\boldsymbol{M}_{h}\|}{\|\boldsymbol{M}\|} 𝒖𝒖h1,h𝒖1\frac{\|\boldsymbol{u}-\boldsymbol{u}_{h}\|_{1,h}}{\|\boldsymbol{u}\|_{1}} pphp\frac{\|p-p_{h}\|}{\|p\|} curl𝑯h\|{\rm curl\,}\boldsymbol{H}_{h}\|_{\infty}
44 0.47390.4739 0.47390.4739 0.47550.4755 0.20660.2066 0.36650.3665 0.0213e120.0213e-12
88 0.24910.2491 0.24910.2491 0.25060.2506 0.10820.1082 0.18030.1803 0.1048e120.1048e-12
1616 0.12620.1262 0.12620.1262 0.12710.1271 0.05510.0551 0.08350.0835 0.6821e120.6821e-12
order 0.95450.9545 0.95450.9545 0.95150.9515 0.95290.9529 1.06711.0671 --
Example 4.5 (A higher order approximation).

In this 3D test, the exact solution of the FHD model (16) is given by

ϕ(x,y,z)=(x2+y2+z2)(x2x)(y2y)(z2z),\displaystyle\phi(x,y,z)=(x^{2}+y^{2}+z^{2})(x^{2}-x)(y^{2}-y)(z^{2}-z),
𝑯(x,y,z)=(2x(x2x)(y2y)(z2z)+(2x1)(x2+y2+z2)(y2y)(z2z)2y(x2x)(y2y)(z2z)+(2y1)(x2+y2+z2)(x2x)(z2z)2z(x2x)(y2y)(z2z)+(2z1)(x2+y2+z2)(x2x)(y2y)),\displaystyle\boldsymbol{H}(x,y,z)=\begin{pmatrix}2x(x^{2}-x)(y^{2}-y)(z^{2}-z)+(2x-1)(x^{2}+y^{2}+z^{2})(y^{2}-y)(z^{2}-z)\\ 2y(x^{2}-x)(y^{2}-y)(z^{2}-z)+(2y-1)(x^{2}+y^{2}+z^{2})(x^{2}-x)(z^{2}-z)\\ 2z(x^{2}-x)(y^{2}-y)(z^{2}-z)+(2z-1)(x^{2}+y^{2}+z^{2})(x^{2}-x)(y^{2}-y)\end{pmatrix},
𝑴(x,y,z)=Ms(coth(γH)1γH)𝑯H,\displaystyle\boldsymbol{M}(x,y,z)=M_{s}\left(\coth(\gamma H)-\frac{1}{\gamma H}\right)\frac{\boldsymbol{H}}{H},
𝒖(x,y,z)=(sin(πy),sin(πz),sin(πx)),\displaystyle\boldsymbol{u}(x,y,z)=\begin{pmatrix}\sin(\pi y),&\sin(\pi z),&\sin(\pi x)\end{pmatrix}^{\intercal},
p(x,y,z)=120x2yz40y3z40yz3,\displaystyle p(x,y,z)=120x^{2}yz-40y^{3}z-40yz^{3},

where the parameters MsM_{s}, γ\gamma, ρ\rho, η\eta and μ0\mu_{0} are all choosen as 11.

We use the conforming quadratic (2\mathbb{P}_{2}) element to discretize the variable ϕ\phi, the first order Nédélec edge element NE1NE_{1} [18, 19] to discretize the variables 𝐇\boldsymbol{H} and 𝐌\boldsymbol{M}, the Taylor-Hood 2\mathbb{P}_{2}-1\mathbb{P}_{1} element to discretize the variables 𝐮\boldsymbol{u} and p~\tilde{p}, and the continuous linear (1\mathbb{P}_{1}) element to discretize the variables ψ\psi and pp. With these settings, we see that the theoretical accuracy of the scheme is 𝒪(h2)\mathcal{O}(h^{2}). We list numerical results in Table 3.

Table 3. Numerical results for Example 4.5.
NN (ϕϕh)ϕ\frac{\|\nabla(\phi-\phi_{h})\|}{\|\nabla\phi\|} 𝑯𝑯hc𝑯c\frac{\|\boldsymbol{H}-\boldsymbol{H}_{h}\|_{c}}{\|\boldsymbol{H}\|_{c}} 𝑴𝑴h𝑴\frac{\|\boldsymbol{M}-\boldsymbol{M}_{h}\|}{\|\boldsymbol{M}\|} (𝒖𝒖h)𝒖\frac{\|\nabla(\boldsymbol{u}-\boldsymbol{u}_{h})\|}{\|\nabla\boldsymbol{u}\|} pphp\frac{\|p-p_{h}\|}{\|p\|} curl𝑯h\|{\rm curl\,}\boldsymbol{H}_{h}\|
44 0.13690.1369 0.13690.1369 0.13690.1369 0.04500.0450 0.05350.0535 0.2970e080.2970e-08
88 0.03720.0372 0.03720.0372 0.03720.0372 0.00810.0081 0.01350.0135 0.7640e080.7640e-08
1616 0.00950.0095 0.00950.0095 0.00950.0095 0.00160.0016 0.00340.0034 1.6060e081.6060e-08
order 1.92451.9245 1.92451.9245 1.92451.9245 2.40692.4069 1.98801.9880 --

From Examples 4.3, 4.4 and 4.5, we have the following observations:

  • From Tables 1 and 2, we see that the errors in H1H^{1} semi-norm for ϕ\phi, 𝑯(curl)\boldsymbol{H}({\rm curl\,}) norm for 𝑯\boldsymbol{H}, L2L^{2} norm for 𝑴\boldsymbol{M} and pp, and discrete H1H^{1} norm for 𝒖\boldsymbol{u}, all have the first (optimal) order rate.

  • From Table 4.5, we see that the errors in H1H^{1} semi-norm for ϕ\phi and 𝒖\boldsymbol{u}, 𝑯(curl)\boldsymbol{H}({\rm curl\,}) norm for 𝑯\boldsymbol{H}, L2L^{2} norm for 𝑴\boldsymbol{M} and pp, all have the second (optimal) order rate.

  • From Tables 1 - 3, we can see that the numerical scheme preserves curl𝑯h=0{\rm curl\,}\boldsymbol{H}_{h}=0 exactly.

References

  • [1] Y. Amirat and K. Hamdache. Global weak solutions to a ferrofluid flow model. Mathematical Methods in the Applied Sciences, 31(2):123–151, 2008.
  • [2] Y. Amirat and K. Hamdache. Strong solutions to the equations of a ferrofluid flow model. Journal of Mathematical Analysis and Applications, 353(1):271–294, 2009.
  • [3] Y. Amirat and K. Hamdache. Unique solvability of equations of motion for ferrofluids. Nonlinear Analysis Theory Methods and Applications, 73(2):471–494, 2010.
  • [4] Y. Amirat, K. Hamdache, and F. Murat. Global weak solutions to equations of motion for magnetic fluids. Journal of Mathematical Fluid Mechanics, 10(3):326–351, 2008.
  • [5] D. N. Arnold, R. S. Falk, and J. Gopalakrishnan. Mixed finite element approximation of the vector laplacian with dirichlet boundary conditions. Mathematical Models and Methods in Applied Sciences, 22(09):1250024, 2012.
  • [6] D. N. Arnold, R. S. Falk, and R. Winther. Finite element exterior calculus, homological techniques, and applications. Acta Numerica, 15:1–155, 2006.
  • [7] D. N. Arnold, R. S. Falk, and R. Winther. Finite element exterior calculus: from hodge theory to numerical stability. Bulletin of the American Mathematical Society, 47(2):281–354, 2010.
  • [8] L. Boccardo, F. Murat, and J.-P. Puel. LL^{\infty} estimate for some nonlinear elliptic partial differential equations and application to an existence result. SIAM J. Math. Anal., 23(2): 326–333, 1992.
  • [9] F. Brezzi and M. Fortin. Mixed and hybrid finite element methods. Springer-Verlag, New York, 1991.
  • [10] L. Chen. ifem: an integrated finite element methods package in matlab. Technical report, University of California at Irvine, 2009.
  • [11] P. Ciarlet. The finite element method for elliptic problems. Amsterdam: North-Holland, 1978.
  • [12] M. Crouzeix and P. A. Raviart. Conforming and nonconforming finite element methods for solving the stationary stokes equations i. Revue française d automatique informatique recherche opérationnelle Mathématique, 7(3), 1973.
  • [13] V. Girault and P. A. Raviart. Finite element methods for Navier-Stokes equations. Springer-Verlag, New York, 1986.
  • [14] V. Girault and P.-A. Raviart. Finite element methods for Navier-Stokes equations: theory and algorithms, volume 5. Springer Science & Business Media, 2012.
  • [15] A. Gil, J. Segura, and N. M. Temme, Numerical methods for special functions, Society for Industrial and Applied Mathematics, 2007.
  • [16] P. Knobloch, S. S, and L. Tobiska. Numerical treatment of a free surface problem in ferrohydrodynamics. Pamm, 10(1):573–574, 2010.
  • [17] O. Lavrova, G. Matthies, T. Mitkova, V. Polevikov, and L. Tobiska. Numerical treatment of free surface problems in ferrohydrodynamics. Journal of Physics Condensed Matter, 18(38):2657–2669, 2006.
  • [18] J.-C. Nédélec. Mixed finite elements in 3\mathbb{R}^{3}. Numerische Mathematik, 35(3):315–341, 1980.
  • [19] J.-C. Nédélec. A new family of mixed finite elements in r3r^{3}. Numerische Mathematik, 50(1):57–81, 1986.
  • [20] R. Nochetto, A. Salgado, and I. Tomas. The equations of ferrohydrodynamics: modeling and numerical methods. Mathematical Models and Methods in Applied Sciences, 26(13), 2015.
  • [21] R. Nochetto, A. Salgado, and I. Tomas. On the Dynamics of Ferrofluids: Global Weak Solutions to the Rosensweig System and Rigorous Convergence to Equilibrium. SIAM Journal on Mathematical Analysis, 51(6), 4245-4286, 2019.
  • [22] Q. Pankhurst, J. Connolly, S. Jones, and J. Dobson. Applications of magnetic nanoparticles in biomedicine. Journal of Physics D-Applied Physics, 36: R167-R181, 2003.
  • [23] R. Rosensweig. Magnetic fluids. Ann. Rev. Fluid Mech., 19: 437-463, 1987.
  • [24] R. Rosensweig. Ferrohydrodynamics. Cambridge University Press, Cambridge, UK, 1985.
  • [25] M. I. Shliomis. Effective viscosity of magnetic suspensions. Sov. Phys. jetp, 34(6):1291–1294, 1972.
  • [26] M. I. Shliomis. Ferrofluids: Magnetically controllable fluids and their applications. Lecture Notes in Physics, 2002.
  • [27] S. M. Snyder, T. Cader, and B. A. Finlayson. Finite element model of magnetoconvection of a ferrofluid. Journal of Magnetism and Magnetic Materials, 262(2):269–279, 2003.
  • [28] E. Sǔli and D. Mayers, An introduction to Numerical Analysis, Cabridge University Press, 2003.
  • [29] Y. Wu and X. Xie, Energy-preserving mixed finite element methods for a ferrofluid flow model, submitted; arXiv: 2206.03129, 2022.
  • [30] G. Yoshikawa, K. Hirata, F. Miyasaka, and O. Yu. Numerical analysis of transitional behavior of ferrofluid employing mps method and fem. IEEE Transaction on Magnetics, 47(5):1370–1373, 2010.
  • [31] M. Zahn. Magnetic fluid and nanoparticle applications to nanotechnology. Journal of Nanoparticle Research, 3(73):73–78, 2001.
  • [32] G. Zhang, X. He, and X. Yang. Decoupled, Linear, and Unconditionally energy stable fully discrete finite element numerical scheme for a two-phase ferrohydrodynamics model. SIAM Journal on Scientific Computing, 43(1): B167-B193, 2021.