Appl. Math. Mech. -Engl. Ed.   2016, Vol. 37 Issue (11): 1479-1500     PDF       
http://dx.doi.org/10.1007/s10483-016-2107-9
Shanghai University
0

Article Information

Jikun ZHAO, Shipeng MAO, Weiying ZHENG
Anisotropic adaptive finite element method for magnetohydrodynamic flow at high Hartmann numbers
Applied Mathematics and Mechanics (English Edition), 2016, 37(11): 1479-1500.
http://dx.doi.org/10.1007/s10483-016-2107-9

Article History

Received Jan. 25, 2016
Revised Jun. 6, 2016
Anisotropic adaptive finite element method for magnetohydrodynamic flow at high Hartmann numbers
Jikun ZHAO1, Shipeng MAO2,3, Weiying ZHENG2,3     
1. School of Mathematics and Statistics, Zhengzhou University, Zhengzhou 450001, China;
2. State Key Laboratory of Scientific and Engineering Computing(LSEC) and Institute of Computational Mathematics, Academy of Mathematics and Systems Science(AMSS), Chinese Academy of Sciences, Beijing 100190, China;
3. School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China
Abstract: This paper presents an anisotropic adaptive finite element method (FEM) to solve the governing equations of steady magnetohydrodynamic (MHD) duct flow.A residual error estimator is presented for the standard FEM,and two-sided bounds on the error independent of the aspect ratio of meshes are provided.Based on the Zienkiewicz-Zhu estimates,a computable anisotropic error indicator and an implement anisotropic adaptive refinement for the MHD problem are derived at different values of the Hartmann number.The most distinguishing feature of the method is that the layer information from some directions is captured well such that the number of mesh vertices is dramatically reduced for a given level of accuracy.Thus,this approach is more suitable for approximating the layer problem at high Hartmann numbers.Numerical results show efficiency of the algorithm.
Key words: magnetohydrodynamic (MHD)flow     posteriori error estimate     anisotropic adaptive finite element method (FEM)    
1 Introduction

The flow problem of viscous, incompressible, electrically conducting fluids in the channels and ducts under a uniform oblique magnetic field is of great interest, because it has many practical applications in the field of magnetohydrodynamics (MHD), such as the blood flow control and measurements, MHD flowmeters, MHD power generation, and accelerators. Only for some very special cases, the problem can be exactly solved. Therefore, for the sake of application, an effective numerical method to solve the MHD flow problem becomes an very important research topic.

It is well-known that, for large values of the Hartmann number, the MHD flow problem is convection-dominated such that the exact solution may display localized phenomena, such as interior or boundary layers. For solving the MHD flow problem, there have been many research works using various numerical methods, such as the finite element, finite difference, and boundary element methods. The readers car refer to Refs. [1]-[2] and many references cited therein. Unfortunately, because of the lack of stability or accuracy, most of conventional numerical methods cannot solve the layer problems efficiently. In Ref. [1], it was pointed out that the common deficiency among the existing numerical methods is that they produce accurate results only in several special configurations for the MHD duct flow problems, but the Hartmann number Ha cannot exceed about 102. However, in many industrial applications, the important range of the Hartmann number is 102 < Ha < 106. For this reason, some seemingly robust numerical methods with large Hartmann numbers were suggested, such as the stabilized finite element method (FEM)[1] and least-squares FEM[2] using residual-free bubble functions. Even so, for large Hartmann numbers, some spurious oscillations still appear near the boundary layer regions when these two bubble-stabilized methods are used (see Refs. [1] and [2] or also the discussion on this in Ref. [3]). Later, a tailored finite point method was proposed[3], and high accuracy was achieved even for large Hartmann numbers. Beyond that, the boundary elements method[4] and the two-level element free Galerkin method[5] were also presented for solving the MHD duct flow problem at high Hartmann numbers, respectively. The theoretical error analysis of these methods above remains seriously poor except for the least-squares FEM[2]. In addition, those methods were only investigated on isotropic meshes where the aspect ratio of mesh elements was uniformly bounded.

For such layer problems, a special anisotropic mesh adaptivity is more desirable, i.e., the mesh elements should adapt in both size and shape in order to approximate the solution better. There have been a well-developed literature on a posteriori error estimation but on isotropic meshes (see Refs. [6] and [7] and the references therein). Some types of the posteriori error estimate have already been extended to anisotropic meshes with large aspect ratio by Kunert[8-10], but no anisotropic mesh refinements were made. Due to large aspect ratio, there appears the alignment measure in a posteriori error estimation to quantify the alignment between the solution and the mesh, which is one of the most important contributions of Kunert.

Based on a priori error estimate, the concept of metrics was mostly used in anisotropic mesh adaptation[11-12]. A priori error analysis on anisotropic meshes can be found in Refs. [13] and [14] and the references therein. Moreover, a posteriori error indicator was used in anisotropic mesh adaptation[15]. The concept of metrics was combined with the solution of a dual problem and a posteriori error analysis in order to generate an adaptive mesh[16]. The metric field was directly obtained from a posteriori error estimate and used in the anisotropic mesh adaptation[17].

In this paper, we present an anisotropic adaptive FEM to solve the MHD duct flow problem in a straight channel of uniform cross-section. By the analytic tools of Kunert[9], a residual error estimator is proposed and provides two-sided bounds on the error on anisotropic meshes. However, the error estimator is not available for anisotropic adaptive refinement, because it cannot represent the contributions to the whole error in different directions, and the alignment measure, in which the exact solution is contained, enters the upper error bound. In order to use the alignment measure as a guide to design a proper anisotropic mesh, we reformulate its definition and use the well-known Zienkiewicz-Zhu (Z-Z) estimates[6]. Then, a computable anisotropic error indicator is derived, and the corresponding locally error indicators are used to adjust the mesh sizes in different directions. Note that an analogous error indicator was proposed for anisotropic adaptive refinement for elliptic and parabolic problems in Ref. [15]. Finally, we implement an adaptive algorithm using the software FREEFEM++[18] with the mesh generator Bamg[19] based on a right metric field obtained by the locally error estimates and Z-Z estimates of the error gradient matrix. By this method, the layer information from some directions can be well captured such that the number of mesh vertices is dramatically reduced for a given level of accuracy. Numerical results confirm the error analysis.

The paper is organized as follows. We first specify our notation and give some preliminary results in Section 2. Section 3 is devoted to analytical tools, such as inverse inequalities, alignment measure, and anisotropic interpolation estimates. An error estimator is proposed for the MHD problem, and the two-sided bounds on the error are derived on anisotropic meshes in Section 4. In Section 5, we derive an anisotropic error indicator and implement an adaptive algorithm. Therein, a series of numerical simulations for three test problems at different Hartmann numbers are carried out to demonstrate effectiveness of the anisotropic adaptive FEM.

2 Preliminaries

In this section, we give some notation and the anisotropic mesh description, present some preliminary results, and finally introduce the model problem and its conforming discretization.

2.1 Notation

For a bounded domain Ω with the Lipschitz continuous boundary ∂Ω, let S be any given open subset of Ω. For the spaces L2(S) and L2(S)2, the integral inner product and the norm are denoted by (·, ·)S and ||·||S, respectively. If S = Ω, the subscript will be omitted. Let |S| be the Lebesgue measure of S, and in particular, |s| is the length of a segment s. Pk(S) is the space of polynomials of order k or less, where k is a given non-negative integer. For simplicity, instead of xcy and c1xyc2x, we use the abbreviated notation x y and x ~ y, where the constants c, c1, and c2 are independent of x, y, the Hartmann number, and the mesh.

By F = {Th}, we denote a family of triangulations Th of Ω, where any two triangles are either disjoint or share a common vertex or edge. Let Eh be the set of all edges of Th, let Ehint be the set of interior, and let Ehext be the set of boundary. We denote by Nh (Nhint, Nhext) the set of all (interior, boundary) vertices of Th. Define for aNh, σEh, and KTh, Ta := {LTh; aL}, Tσ := {LTh; σ ⊂ L}, and TK := {LTh; LK ≠ ∅}, respectively. Furthermore, three auxiliary subdomains need to be defined by

(1)

Next, the notation nK always denotes the exterior unit normal vector for any given KTh, and nσ denotes the unit normal vector for any given edge σEh, of which the orientation is arbitrarily chosen but coinciding with the exterior normal of the domain Ω for boundary edges and fixed for interior edges.

For a function ϕ and an edge σEhint shared by two triangles K and L, where nσ points from K to L, we define the jump operator through σ by

(2)

For any σEhext, set .

2.2 Anisotropic meshes

For an arbitrary (anisotropic) triangle KTh, we enumerate its vertices such that P0P1 is the longest edge and |P1P2| ≥ |P0P2|. Moreover, we define two orthogonal vectors pi with the length hi, K := |pi| (i = 1, 2), as shown in Fig. 1. Notice that h1, Kh2, K, set hmin, K := h2, K, and hmax, K := h1, K. For σ ⊂ ∂K, let

Fig. 1 Triangle K
(3)

Define two 2 × 2 matrices AK and CK by

(4)

Additionally, we require that the mesh Th satisfies the following requirements:

(i) The number of triangles contained in Ta is bounded uniformly for each fixed vertex aNh.

(ii) The dimension of adjacent triangles is close to each other, i.e.,

(5)

For convenience, for the common edge σ shared by two triangles K and L, set

(6)

An advantage of doing this is that they are no longer related to either of K or L but only to σ. For the boundary edges, the definitions are changed in the obvious way. Obviously, they satisfy hσ ~ hσ, K ~ hσ, L and hmin, σ ~ hmin, K ~ hmin, L. For more details on anisotropic meshes, see Refs. [8] and [9].

2.3 Definition of problem and its conforming discretization

In a straight channel of a uniform cross-section Ω, we assume that there exists a laminar fully developed flow of a viscous, incompressible, and electrically conducting fluid. Let u be the velocity, and let B be the induced magnetic field. Here, the direction of the uniform transverse applied magnetic field B0 may be arbitrary to the x-axis, and the fields u and B are parallel to the z-axis. Further, The fluid is driven down by a constant pressure gradient. The governing equations for the above duct flow in a dimensionless form with suitable boundary conditions can be posed as follows[4, 20]:

(1)

where Ha = B0l(δ/ν)1/2 is the Hartmann number, l is the characteristic length of the duct, B0 is the intensity of the external magnetic field, and ν and δ are the viscosity coefficient and electric conductivity of the fluid, respectively. a = (cos α, sin α)T, where α is the angle between the x-axis and the externally applied magnetic field B0. ∂Ω = ΓD ∪ ΓN, where ΓD ∩ ΓN = ∅ and ΓD has a positive measure. We call ΓN the conducting part and ΓD the insulated part of the boundary ∂Ω.

Let W0B(Ω) := H01(Ω) × HB1 (Ω), where two particular subspaces of H1(Ω) are

(7)

Due to the Poincaré inequality, the space W0B(Ω) will be equipped with the product norm,

(8)

The weak form of the problem (1) is to find (u, B)∈W0B(Ω) such that

(2)

In order to define the finite element approximation, let Vh(Ω) be a subspace of H1(Ω) consisting of continuous piecewise affine functions on the mesh Th, and

(9)

Then, the standard FEM is to find (uh, Bh)∈W0Bh (Ω) such that

(3)

The existence and uniqueness of solutions to (2) and (3) can be easily obtained by the wellknown Lax-Milgram lemma. A priori error analysis can be found in Ref. [21].

3 Analytical tools

In order to treat anisotropic elements, some analytical tools in the anisotropic setting have to be introduced here, which are taken from Refs. [8] and [9].

3.1 Inverse inequalities

Inverse inequalities for bubble functions are very important in deriving lower error bounds. As usual, we first introduce the bubble functions[7]. For an arbitrary triangle K, denote the barycentric coordinates by λK, 1, λK, 2, and λK, 3. We define the element bubble function bK by

(10)

Let σ be an inner edge shared by K1 and K2. We here enumerate all the vertices of K1 and K2 such that the two vertices of σ are first numbered. We define the edge bubble function bσ by

(11)

Obvious modifications need to be done for boundary edges. For simplicity, bK and bσ are assumed to be extended by zero outside their original domain of definition.

For a given edge σ of a triangle K, an extension operator Fext: P0(σ) → P0(K) is defined as follows:

(12)

By standard scaling arguments, we can easily derive the following anisotropic inverse inequalities.

Lemma 1 Assume that ϕKP0(K) and ϕσP0(σ), where σ∂K. Then,

(13)
3.2 Alignment measure and anisotropic interpolation estimates

From a heuristic point of view, if the (directional) derivative of the solution displays little change in some direction, in that direction, we should stretch the mesh elements. The better the alignment is between the anisotropic mesh and the anisotropy of solution, the more accurate the error estimates can be. All practical applications intuitively follow this concept. Next, we should introduce the alignment measure m1(v, Th) in order to quantify this alignment.

For vH1(Ω) and a family of triangulations {Th}, the alignment measure m1: H1(Ω) × {Th} → R is defined by

(4)

The alignment measure has the following property:

(14)

The above property implies that, if the mesh Th is well aligned with an anisotropic function v, it leads to a small alignment measure. In practice, for sensible anisotropic meshes, one almost always obtains m1(v, Th) ~ 1. In this paper, we should apply this property to anisotropic mesh adaption (see numerical experiments in Section 5).

Consider a node aNh. The local L2-projection Pa: H1(ωa) → P0(ωa) is uniquely defined by . Then, the Clément interpolation operator I0: H01(Ω) → Vh(Ω) ∩ H01(Ω) is defined by

(15)

and IB: HB1 (Ω) → Vh(Ω) ∩ HB1 (Ω) by

(16)

where λa is the (piecewise linear) basis function related to the vertex a, and NNext := {aNhext; a∈ΓN}.

Finally, we state the anisotropic interpolation error estimates based on the alignment measure.

Lemma 2 For all functions vH01(Ω) and QHB1 (Ω), there hold

(5)
(6)
(7)
(8)

where ENext := {σEhext; σ ⊂ ΓN}.

4 Residual error estimation

In this section, we present the residual error estimator for the conforming approximation (3) and show that it provides two-sided bounds on the error ||(uuh, BBh)||W on anisotropic meshes.

Let (uh, Bh) be the conforming finite element approximation. Here, we define the element residuals over an element K by

(17)

and edge residuals by

(18)

Obviously, R1, K = 1+Ha a·▽Bh and R2, K = Ha a·▽uh hold for piecewise linear functions as considered here.

The local residual error estimators η1, K and η2, K for an element K are defined by

(9)
(10)

and the global terms are

(19)

separately.

Theorem 3 Let (u, B) be the weak solution of problem (1), and let (uh, Bh) be the corresponding finite element approximation defined by (3). Then, the error is bounded globally from above by

(11)

Proof For the convenience of expression, we set eu = uuh and eB = BBh. Since (a · ▽eB, eu) = −(a · ▽eu, eB), due to integration by parts, the Galerkin orthogonality and Lemma 2 yield

(20)

which concludes the proof.

Theorem 4 Under the assumption of Theorem 3, the error is bounded locally from below by

(12)
(13)

for all KTh.

Proof The proof of lower error bounds needs to use the bubble functions and corresponding anisotropic inverse inequalities (see Lemma 1). Here, we only need to prove the lower bound (12). The remaining one can be derived analogously.

Observe R1, K = 1 + Δuh + Ha a · ▽BhP0(K) and set

(21)

The Cauchy-Schwarz inequality and integration by parts yield

(14)

The inverse inequalities of Lemma 1 lead to

(22)

which, together with (14), result in

(15)

For an inner edge σEhint, set

(23)

Integration by parts and (2) for (v, Q) = (wσ, 0) implies that

(16)

Due to Lemma 1, we have the following relations:

(24)

which, together with (15) and (16), lead to

(17)

For an edge on the Dirichlet boundary, nothing needs to be done because R1, σ = 0. Hence, combining (15) and (17), we conclude the lower error bound (12).

Remark 1 From Theorems 3 and 4, the lower bounds contain additional L2-error terms Ha||uuh||ωK and Ha||BBh||ωK that do not appear in the upper error bound (11). Therefore, the two-sided bounds on the error do not correspond completely. We point out that a very similar situation can be found for error estimators for the convection-diffusion problem[10, 22-23]. Moreover, only if the L2-error terms is dominated by the H1-error terms ||▽(uuh)||ωK and ||▽(BBh)||ωK, the upper and lower bounds on the error will be of the same quality. Hence, the additional L2-error terms are mainly due to the H1-error terms, which is confirmed by our numerical experiments in the next section.

5 Numerical experiments 5.1 Anisotropic error indicator

Due to the large aspect ratio of anisotropic mesh, the alignment measure m1(·, Th) enters the upper error bound in order to measure how well the alignment between the solution and the mesh does in the sense: the worse the alignment is, the larger the alignment measure is. In other words, a sufficiently good mesh makes sure that the alignment measure is O(1), i.e.,

(18)

In turn, the relation (18) can be used as a guide to design a proper anisotropic mesh. To this end, the definition (4) of the alignment measure m1(v, Th) is reformulated by

(19)

where r1 and r2 are the corresponding unit vectors of p1 and p2 on each KTh, respectively, and so CK = (p1, p2) = (hmax, Kr1, hmin, Kr2). Therefore, the relation (18) holds, provided that

(20)

This formulation can be used to indicate the direction of adaptive mesh refinement. More specifically, due to (19), the upper error bound (11) holds as long as we have the following estimate:

(21)

where

(25)

The relation (20) implies η(1) ~ η(2). However, the estimate (21) is not available, because the exact solution (u, B) is usually unknown. To overcome this, we use the well-known Z-Z estimate[6]. The gradient (▽u, ▽B) of the exact solution can be replaced by a recovered gradient (▽Ru, ▽RB) by local averaging such as an approximate L2-projection of (▽uh, ▽Bh) onto Vh × Vh, i.e.,

(26)

Numerical experiments in Ref. [24] showed that the averaging technique was quite more reliable on anisotropic meshes than expected. After this replacement, we introduce the global anisotropic error indicator and the local anisotropic error indicator by

(27)

In conclusion, the relation (18) implies that the adaptive algorithm should be designed such that the local anisotropic error indicators and satisfy, on each triangle K,

(28)
5.2 Adaptive algorithm

We should implement an adaptive algorithm using the software FREEFEM++[18] with the mesh generator Bamg[19] based on a right metric field which will be built later. The goal of this adaptive algorithm is to generate a suitable anisotropic mesh so that the relative estimated error is close to a preset tolerance ϕTOL, i.e.,

(29)

A sufficient condition to generate such an anisotropic mesh is to make sure that for all KTh, it holds

(22)

where |Th| denotes the number of elements in Th, as well as the number |Nh| of mesh vertices which will be used later. In order to get data at the mesh vertices, at each vertex a, we introduce the anisotropic error indicator defined by

(30)

Since

(31)

there holds (22), whenever we generate a mesh satisfying, for all vertices aNh,

(32)

With the mesh generator Bamg, the metric field M is needed to be given at each vertex a as continuous P1 finite element functions, namely, the two eigenvalues λ1, a, λ2, a (λ1, aλ2, a) and the corresponding eigenvectors r1, a, r2, a such that M = RTΛR, where

(33)

Then, in a vicinity of the vertex a, the wanted mesh size can be defined by the metric field M such that the size h is equal to |x|√xTM(a)x in the direction x∈R2, where |x| := √xTx.

Then, the adaptive algorithm is designed. First, the values and at all the vertices a of the mesh are defined by

(34)

Thus, the error at each vertex a in the direction of maximum stretching can be represented by , whereas the error at each vertex a in the direction of minimum stretching is denoted by . Second, we compute the value of mesh size h1, a (h2, a) for all vertices aNh by averaging all the values hmax, K (hmin, K) corresponding to the neighboring triangles K in Ta. The algorithm is carried out as follows. For i = 1, 2, if

(35)

then the value of λi, a is set to . If

(36)

then the value of λi, a is set to be . Otherwise, λi, a is set to be hi, a−1. Finally, for all vertices aNh, let r1, a and r2, a be, respectively, the unit eigenvectors corresponding to the smallest and largest eigenvalues of the Z-Z estimates of the error gradient matrix defined by an average

(37)

With the resulting metric M, we build a new anisotropic mesh using Bamg, which should insure m1R (uuh, Th) ~ 1 and m1R (BBh, Th) ~ 1, where m1R (uuh, Th) and m1R (BBh, Th) are the Z-Z estimates of m1(uuh, Th) and m1(BBh, Th) obtained by replacing the gradient (▽u, ▽B) of the exact solution by the recovered one (▽Ru, ▽RB), i.e.,

(38)

respectively. According to the numerical results of the next subsection, it always holds that both m1R (uuh, Th) and m1R (BBh, Th) are close to 1, which indicates that the anisotropy of meshes is well adapted to the exact solution of the problem. Therefore, an efficient error estimation is to be expected. In the future, we plan to extend such an efficient anisotropic adaptive FEM to other problems, e.g., nonconforming FEMs[25-26] for fourth-order elliptic singular perturbation problems with boundary layers[27].

5.3 Numerical results

In this subsection, for the three test problems, a series of numerical simulations are performed by the anisotropic adaptive FEM developed in this paper. In all the test problems, we consider the domain Ω = (0, 1) × (0, 1) and set the tolerance ϕTOL = 0.01.

Example 1 (The Shercliff problem) For this example, there exists an analytic solution, and the numerical data are available in Refs.[1] and [2] such that a comparison can be done between the approximate solution and the exact solution. Here, the walls of the channel are insulated (B = 0 on ∂Ω), and on the solid walls, the velocity is zero (u = 0 on ∂Ω). The external magnetic field B0 is applied in the x-axis (α = 0) (see Fig. 2).

Fig. 2 Boundary conditions of Shercliff problem

For the Hartmann numbers Ha = 100 and Ha = 500, the comparisons are done between the approximate solution (uh, Bh) and the exact solution (u, B) in Tables 1 and 2, respectively. One can see that two results are comparable. In addition, Table 3 presents the corresponding mesh information and Z-Z estimates of the alignment measures after some iterations during the adaptation process. For the sake of comparison, the corresponding meshes and contour plots for approximate solutions are shown in Figs. 3-4 for Ha = 100 and Ha = 500, respectively. It is worth mentioning that the use of anisotropic adaptive FEM can capture the boundary or interior layers in some directions and allow the number of vertices dramatically reduced for a given level of accuracy. In the next two examples, the superior performance will be further demonstrated for much larger Hartmann numbers, such as 103, 104, and 105.

Table 1 Approximate and exact solutions for Shercliff problem at Ha = 100
Table 2 Approximate and exact solutions for Shercliff problem at Ha = 500
Table 3 Mesh information and Z-Z estimates of alignment measures for Shercliff problem
Fig. 3 Corresponding meshes for Shercliff problem at Ha = 100 (left) and Ha = 500 (right)
Fig. 4 Approximate velocity fields for Shercliff problem at Ha = 100 (left) and Ha = 500 (right)
Fig. 5 Approximate magnetic fields for Shercliff problem at Ha = 100 (left) and Ha = 500 (right)

Example 2 (The two-dimensional square-channel flow with an oblique applied magnetic field) The same MHD problem as Example 1 is considered, except that the externally applied magnetic field B0 has various positive angles α with the x-axis (see Fig. 6). Here, for the values of α = 0, π/4, and π/3, numerical computations are carried out at different Hartmann numbers. The resulting meshes and contour plots for approximate solutions generated by the anisotropic adaptive FEM are presented at Hartmann numbers Ha = 103, 104, and 105, respectively, in Figs. 7-15. The numerical results show that when Hartmann number increases, the boundary or interior layers are well dealt with by anisotropic meshes. The corresponding mesh information and approximate alignment measures are presented in Tables 4-6.

Fig. 6 Boundary conditions of Example 2
Fig. 7 Corresponding meshes for Example 2 at Ha = 103 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 8 Approximate velocity fields for Example 2 at M = 103 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 9 Approximate magnetic fields for Example 2 at M = 103 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 10 Corresponding meshes for Example 2 at Ha = 104 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 11 Approximate velocity fields for Example 2 at Ha = 104 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 12 Approximate magnetic fields for Example 2 at Ha = 104 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 13 Corresponding meshes for Example 2 at Ha = 105 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 14 Approximate velocity fields for Example 2 at Ha = 105 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Fig. 15 Approximate magnetic fields for Example 2 at Ha = 105 with α = 0 (left), α = π/4 (middle), and α = π/3 (right)
Table 4 Mesh information and Z-Z estimates of alignment measures for Example 2 at Ha = 103
Table 5 Mesh information and Z-Z estimates of alignment measures for Example 2 at Ha = 104
Table 6 Mesh information and Z-Z estimates of alignment measures for Example 2 at Ha = 105

Example 3 (The two-dimensional square-channel flow with a partly conducting boundary) Here, the external magnetic field B0 is perpendicular to the wall at x = −1 (i.e., α = 0). For a length 2l at the center, the wall is electrically insulated (see Fig. 16). In Figs. 17-25, the anisotropic meshes and contour plots for approximate solutions are presented for l = 0.2, 0.5, and 0.7 at different Hartmann numbers (Ha = 103, 104, and 105), respectively. The corresponding mesh information and approximate alignment measures are presented in Tables 7-9.

Fig. 16 Boundary conditions of Example 3
Fig. 17 Corresponding meshes for Example 3 at Ha = 103 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 18 Approximate velocity fields for Example 3 at Ha = 103 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 19 Approximate magnetic fields for Example 3 at Ha = 103 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 20 Corresponding meshes for Example 3 at Ha = 104 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 21 Approximate velocity fields for Example 3 at Ha = 104 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 22 Approximate magnetic fields for Example 3 at Ha = 104 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 23 Corresponding meshes for Example 3 at Ha = 105 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 24 Approximate velocity fields for Example 3 at Ha = 105 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Fig. 25 Approximate magnetic fields for Example 3 at Ha = 105 with l = 0.2 (left), l = 0.5 (middle), and l = 0.7 (right)
Table 7 Mesh information and Z-Z estimates of alignment measures for Example 3 at Ha = 103
Table 8 Mesh information and Z-Z estimates of alignment measures for Example 3 at Ha = 104
Table 9 Mesh information and Z-Z estimates of alignment measures for Example 3 at Ha = 105
6 Conclusions

In this paper, an anisotropic adaptive FEM is proposed to solve the MHD duct flow at high Hartmann numbers. The most distinguish feature of this method is that the layer information from some directions is captured well such that the number of mesh vertices is dramatically reduced for a given level of accuracy. Further, an adaptive algorithm is implemented for several examples on a rectangular domain, where it is shown that even for high values of the Hartmann number, this method gives accurate and stable results. Finally, we note that the anisotropic adaptive FEM can be applied to channels with arbitrary cross-sections.

References
[1] Nesliturk, A. I., & Tezer-Sezgin, M. The finite element method for MHD flow at high Hartmann numbers. Comput. Methods Appl. Mech. Engrg., 194(9-11), 1201-1224 (2005) doi:10.1016/j.cma.2004.06.035
[2] Hsieh, P. W., & Yang, S. Y. A bubble-stabilized least-squares finite element method for steady MHD duct flow problems at high Hartmann numbers. J. Comput. Phys., 228(22), 8301-8320 (2009) doi:10.1016/j.jcp.2009.08.007
[3] Hsieh, P. W., Shih, Y., & Yang, S. Y. A tailored finite point method for solving steady MHD duct flow problems with boundary layers. Commun. Comput. Phys., 10(1), 161-182 (2011) doi:10.4208/cicp.070110.020710a
[4] Hosseinzadeh, H., Dehghan, M., & Mirzaei, D. The boundary elements method for magnetohydrodynamic (MHD) channel flows at high Hartmann numbers. Appl. Math. Model., 37(4), 2337-2351 (2013) doi:10.1016/j.apm.2012.05.020
[5] Zhang, L., Ouyang, J., & Zhang, X. The two-level element free Galerkin method for MHD flow at high Hartmann numbers. Phys. Lett. A, 372(35), 5625-5638 (2008) doi:10.1016/j.physleta.2008.05.088
[6] Ainsworth, M., & Oden, J. T. A Posteriori Error Estimation in Finite Element Analysis. John Wiley & Sons, New York (2000)
[7] Verfürth, R. A Review of a Posteriori Error Estimation and Adaptive Mesh-Refinement Techniques, Teubner-Wiley, Stuttgart (1996)
[8] Kunert, G. A Posteriori Error Estimation for Anisotropic Tetrahedral and Triangular Finite Element Meshes, Ph. D. dissertation, Chemnitz Univesity of Technology (1999)
[9] Kunert, G. A posteriori residual error estimator for the finite element method on anisotropic tetrahedral meshes. Numer. Math., 86(3), 471-490 (2000) doi:10.1007/s002110000170
[10] Kunert, G. A posteriori error estimation for convection dominated problems on anisotropic meshes. Math. Method. Appl. Sci., 26(7), 589-617 (2003) doi:10.1002/(ISSN)1099-1476
[11] Frey, P. J., & Alauzet, F. Anisotropic mesh adaptation for CFD computations. Comput. Methods Appl. Mech. Engrg., 194(48-49), 5068-5082 (2005) doi:10.1016/j.cma.2004.11.025
[12] Dolejšı, V. Anisotropic mesh adaptation for finite volume and finite element methods on triangular meshes. Comput. Visual. Sci., 1(3), 165-178 (1998) doi:10.1007/s007910050015
[13] Apel, T. Anisotropic Finite Elements:Local Estimates and Applications, Teubner, Stuttgart (1999)
[14] Chen, S., & Xiao, L. Interpolation theory of anisotropic finite elements and applications. Sci. China Ser. A, 51(8), 1361-1375 (2008) doi:10.1007/s11425-008-0107-y
[15] Picasso, M. An anisotropic error indicator based on Zienkiewicz-Zhu error estimator:application to elliptic and parabolic problems. SIAM J. Sci. Comput., 24(4), 1328-1355 (2003) doi:10.1137/S1064827501398578
[16] Formaggia, L., Micheletti, S., & Perotto, S. Anisotropic mesh adaption in computational fluid dynamics:application to the advection-diffusion-reaction and the Stokes problems. Appl. Numer. Math., 51(4), 511-533 (2004) doi:10.1016/j.apnum.2004.06.007
[17] Kuate, R. Anisotropic metrics for finite element meshes using a posteriori error estimates:Poisson and Stokes equations. Eng. Comput.-Germany, 29(4), 497-505 (2013)
[18] Hecht, F., Pironneau, O., Morice, J., Le Hyaric, A., and Ohtsuka, K. Freefem++documentation, version 3.19-1. http://www.freefem.org/ff++ (2012)
[19] Hecht, F. Bamg:bidimensional anisotropic mesh generator. http://www-rocq1.inria.fr/gamma/cdrom/www/bamg/eng.htm (1998)
[20] Scandiuzzi, R., & Schrefler, B. FEM in steady MHD duct flow problems. Int. J. Numer. Meth. Engrg., 30(4), 647-659 (1990) doi:10.1002/(ISSN)1097-0207
[21] Meir, A. Finite element analysis of magnetohydrodynamic pipe flow. Appl. Math. Comput., 57(2), 177-196 (1993)
[22] Kay, D. The reliability of local error estimators for convection-diffusion equations. IMA J. Numer. Anal., 21(1), 107-122 (2001) doi:10.1093/imanum/21.1.107
[23] Verfürth, R. A posteriori error estimators for convection-diffusion equations. Numer. Math., 80(4), 641-663 (1998) doi:10.1007/s002110050381
[24] Babuška, I., Strouboulis, T., Upadhyay, C. S., Gangaraj, S. K., & Copps, K. Validation of a posteriori error estimators by numerical approach. Int. J. Numer. Meth. Engrg., 37(7), 1073-1123 (1994) doi:10.1002/(ISSN)1097-0207
[25] Chen, H., Chen, S., & Qiao, Z. C0-nonconforming tetrahedral and cuboid elements for the three-dimensional fourth order elliptic problem. Numer. Math., 124(1), 99-119 (2013) doi:10.1007/s00211-012-0508-2
[26] Chen, H. R., Chen, S. C., & Qiao, Z. H. C0-nonconforming triangular prism elements for the three-dimensional fourth order elliptic problem. J. Sci. Comput., 55(3), 645-658 (2013) doi:10.1007/s10915-012-9652-1
[27] Chen, H., & Chen, S. Uniformly convergent nonconforming element for 3-D fourth order elliptic singular perturbation problem. J. Comput. Math., 32(6), 687-695 (2014) doi:10.4208/jcm