Taylor-Hood like finite elements for nearly incompressible strain gradient elasticity problems
Abstract.
We propose a family of mixed finite elements that are robust for the nearly incompressible strain gradient model, which is a fourth-order singular perturbed elliptic system. The element is similar to [C. Taylor and P. Hood, Comput. & Fluids, 1(1973), 73-100] in the Stokes flow. Using a uniform discrete B-B inequality for the mixed finite element pairs, we show the optimal rate of convergence that is robust in the incompressible limit. By a new regularity result that is uniform in both the materials parameter and the incompressibility, we prove the method converges with order to the solution with strong boundary layer effects. Moreover, we estimate the convergence rate of the numerical solution to the unperturbed second-order elliptic system. Numerical results for both smooth solutions and the solutions with sharp layers confirm the theoretical prediction.
Key words and phrases:
Mixed finite elements; Nearly incompressible strain gradient elasticity; Uniform error estimate2020 Mathematics Subject Classification:
Primary 65N15, 65N30; Secondary 74K201. Introduction
The strain gradient models have drawn great attention recently because they capture the size effect of nano-materials for plasticity [FleckHutchinson:1997] as well as for mechanical meta-materials [Khakalo:2020] by incorporating the higher-order strain gradient and the intrinsic material length scale into the constitutive relations. Studies from the perspective of modeling, mechanics and mathematics may date back to 1960s [Koiter:19641, Mindlin:1964, Hlavacek:19692], while large-scale simulations are relatively recent [zervos:20092, Zybell:2012, Rud:2014, Phunpeng:2015]. Different methods such as H2 conforming finite element methods [zervos:20091, Fisher:2010], H1 conforming mixed finite element methods [Aravas:2002, Phunpeng:2015], nonconforming finite element methods [LiMingShi:2017, LiaoM:2019, LiMingWang:2021], discontinuous Galerkin methods [Engel:2002], isogeometric analysis [Klassen:2011, Niiranen:2016], and meshless methods [Askes:2002] have been used to simulate the strain gradient elastic models with different complexity, just to mention a few. One of the numerical difficulties is that the number of the materials parameters is too large [Mindlin:1964], another is that the materials parameters may cause boundary layer or numerical instability when they tend to certain critical values [Engel:2002].
The strain gradient elasticity model proposed by Altan and Aifantis [Altan:1992] seems the simplest one among them because it contains only one material parameter besides the Lamé constants, while it still models the size effect adequately [Aifantis:1999]. This model is described by the following boundary value problem:
(1.1) |
where is a smooth domain, is the displacement, is the normal derivative of , and are the Lamé constants, and is the microscopic parameter such that , which stands for the intrinsic length scale. Besides modeling the strain gradient elasticity, the system (1.1) may also be regarded as a vector analog of the fourth-order singular perturbed problem, which usually models a clamped plate problem [John:1973, Schuss:1976, Semper:1992, Semper:1994, Tai:2001, Brenner:2011], or arises from a fourth-order perturbation of the fully nonlinear Monge-Ampère equation [Brenner:2011b, Feng:2014]. System of the form (1.1) may also come from the linearized model in MEMS [Laurencot:2017].
In the present work, we are interested in (1.1) for the nearly incompressible materials. Such materials are commonly used in industry and a typical example is natural rubber. To the best of our knowledge, the studies on the approximation of incompressible and nearly incompressible strain gradient elasticity have not been sufficiently addressed in the literature, although vast efforts have been devoted to finite element approximation of the incompressible and nearly incompressible elasticity problems; See e.g., [Hermann:1965, Vog:1983, Simo:1985, Babuska:1992, Brenner:1992, BraessMing:2005, Auricchio:2013]. In [FleckHutchinson:1997]*§III. C, the authors studied the incompressible limit of the strain gradient model. Mixed finite elements for the incompressible Fleck-Hutchinson strain gradient model have been designed and tested in [ShuKingFleck:1999]. A finite element method has been tested for the nearly incompressible strain gradient model in [Wei:2006]. A mixed finite element, which approximated the displacement with Bogner-Fox-Schmidt element [BFS:1965] and approximated the pressure with the node quadrilateral element, was constructed for the coupled stress model in [Fisher:2011], and bore certain similarities with problem (1.1). Recently, Hu and Tian [Tian:2021] have proposed several robust elements for the two-dimensional strain gradient model in the framework of reduced integration. Unfortunately, none of the above work proved the robustness of the proposed elements rigorously in the incompressible limit.
Following the classical approach dealing with the nearly incompressible elasticity problem [Hermann:1965], we introduce an auxiliary variable “pressure” and recast (1.1) into a displacement-pressure mixed variational problem, i.e., the so-called formulation. We approximate the displacement by augmenting the finite element space in [GuzmanLeykekhmanNeilan:2012] with certain new bubble functions. The original motivation for the bubble functions is to design the stable finite element pair for the Stokes problem [Arnold:1984, Bernardi:1985]. The augmented bubble functions help out in dealing with the extra constraints such as the divergence stability in Stokes problem and the high order consistency error [Tai:2001, GuzmanLeykekhmanNeilan:2012, Zhang:2012]. Such idea has been exploited by one of the authors to design robust finite elements for the strain gradient elasticity model [LiMingShi:2017]. Besides, we employ the standard continuous Lagrangian finite element of one order lower than that for the displacement to approximate the pressure. Such a finite element pair is similar to the Taylor-Hood element in the Stokes flows [Hood:1973] which is scheme and continuous pressure approximation. For both smooth solutions and solutions with strong boundary layer effects, these mixed finite element pairs are robust in the incompressible limit, here the robustness is understood in the sense that the rate of convergence is uniform in both and . The bubble function spaces in approximating the displacement are defined by certain orthogonal constraints, and the explicit representations of these spaces are desired for the sake of implementation. We achieve this with the aid of the Jacobi polynomial. In addition to perspicuous results in view of analytics, such representation lends itself to the construction of the analytical shape functions for the approximating space of the displacement. Though we focus on the two-dimensional problem, the element may be readily extended to the three-dimensional problem. cf., Remark 3.3.
By standard mixed finite element theory [BBF:2013], a discrete B-B inequality that is uniform in is needed for the well-posedness of the mixed discretization problem. This B-B inequality reduces to the remarkable B-B inequality for the Stokes problem when tends to zero. A natural way to prove the discrete B-B inequality is to construct a uniformly stable Fortin operator [Fortin:1977, Winther:2013, MardalTaiWinther:2002], which does not seem easy due to the complication of the constraints. To this end, we construct a quasi-Fortin operator that takes different forms for small as well as large . This quasi-Fortin operator is bounded in a weighted energy norm in the corresponding regimes of . Besides the discrete B-B inequality, another ingredient in proving the robustness is a new regularity result for the solution of (1.1) that is uniform in both and , which is crucial to prove the convergence rate for the layered solution. The proof combines the method dealing with the nearly incompressible linear elasticity [Vog:1983] and the regularity estimate for the fourth-order singular perturbed problem [Tai:2001, LiMingWang:2021].
The outline of the paper is as follows. In §2, we introduce Altan and Aifantis’ strain gradient model and its mixed variational formulation. We demonstrate the numerical difficulty caused by large , and prove the uniform regularity estimate for problem (1.1). In §3, we construct a family of nonconforming finite elements, and derive the explicit formulations for the bubble spaces. In §4, we use the nonconforming elements proposed in §3 together with the continuous Lagrangian finite elements to discretize the mixed variational problem and prove the optimal rate of convergence. In the last section, we report the numerical results, which are consistent with the theoretical prediction.
Throughout this paper, the constant may differ from line to line, while it is independent of the mesh size , the materials parameter and the Lamé constant .
2. The Mixed Variational Formulation and Regularity Estimates
First we fix some notations. The space of the square-integrable functions defined on a smooth domain is equipped with the inner product and the norm , while is the subspace of with mean value zero. Let be the standard Sobolev space [AdamsFournier:2003] with the norm , while is the closure in of . We may drop in when no confusion may occur. For any vector-valued function , its gradient is a matrix-valued function with components , and the symmetric part of is defined by The divergence operator is defined as . The Sobolev spaces , and of a vector-valued function may be defined similarly as their scalar counterpart. This rule equally applies to the inner products and the norms. The double inner product between tensors equals .
We recast (1.1) into a variational problem: Find such that
(2.1) |
where and the fourth-order tensor and the sixth-order tensor are defined as
respectively. Here is the Kronecker delta function. The strain gradient is a third-order tensor that is defined by .
We are interested in the regime when , which means that the material is nearly incompressible. Proceeding along the same line that leads to [LiMingWang:2021]*Theorem 5, the tensor product of the element (NTW) proposed in [Tai:2001] may be used to approximate (1.1), and the error estimate reads as
where , and is independent of the mesh size , and and . Therefore, the error bound degenerates when is large, and the NTW element does not seem a good candidate for the nearly incompressible material. The following numerical example confirms this observation.
Example 2.1.
Let , and with
It is clear that , hence the material is completely incompressible. The details of the numerical experiment such as the mesh generation, are the same as those in § 5. The relative error in Table 1 shows that the rate of convergence deteriorates when is large.
1/8 | 1/16 | 1/32 | 1/64 | |
1e+00 | 2.681e-01 | 1.373e-01 | 6.698e-02 | 3.334e-02 |
rate | 0.97 | 1.04 | 1.01 | |
1e-06 | 4.550e-02 | 1.244e-02 | 3.001e-03 | 7.467e-04 |
rate | 1.87 | 2.05 | 2.01 | |
1e+00 | 9.995e-01 | 9.979e-01 | 9.916e-01 | 9.682e-01 |
rate | 0.00 | 0.01 | 0.03 | |
1e-06 | 9.752e-01 | 7.233e-01 | 2.502e-01 | 6.561e-02 |
rate | 0.43 | 1.53 | 1.93 |
2.1. The mixed variational formulation
We introduce an auxiliary variable and . We write Problem (2.1) into a mixed variational problem as
(2.2) |
where
It is convenient to define the weighted norm for all as
is a norm over for any and any finite . By Poincaré’s inequality, is a norm over for any . To study the well-posedness of Problem (2.2), we start with the following B-B inequality that is uniform for any .
Lemma 2.2.
For any , there exists such that
(2.3) |
where only depends on but is independent of .
Proof.
By [Galdi:2011]*Theorem III 3.3 and [Costabel:2010]*Proposition 4.1, for any , there exists such that and
(2.4) |
where the constant only depends on .
Because the mean of is zero for any , by Poincaré’s inequality, there exists such that
Lemma 2.3.
There exists a unique and satisfying (2.2), and there exists independent of and such that
(2.5) |
Proof.
By the first Korn’s inequality [Korn:1908, Korn:1909],
(2.6) |
and the H2 Korn’s inequality [LiMingWang:2021]*Theorem 1,
(2.7) |
we obtain
Using (2.3), for any , there exists such that and . This implies
By the standard regularity theory for the elliptic system, we find and provided that , while we are interested in whether the shift estimate is true with a independent constant , this is the objective of the next part.
2.2. Regularity estimates
We aim to study the regularity of the solution of (1.1). Letting , we find satisfying
(2.8) |
in the sense of distribution, where . The Herror for will be given in Theorem 2.9, which is crucial for the regularity estimate of problem (1.1). We reshape (2.8) into a variational problem: Find such that
(2.9) |
By [BacutaBramble:2003], we have the following shift estimate for : There exists independent of such that
(2.10) |
Next we study an auxiliary boundary value problem:
(2.11) |
The a-priori estimate for the solution of the above boundary value problem reads as
Lemma 2.4.
There exists a unique satisfying (2.11), and there exists independent of such that
(2.12) |
Proof.
For any , by the HKorn’s inequality (2.7) and Poincaré’s inequality, there exists such that
The existence and uniqueness of follow from the Lax-Milgram theorem, and
(2.13) |
Now we turn to prove the regularity estimate of problem (2.11). We consider an auxiliary elliptic system: For any and , find and such that the following boundary value problem is valid in the sense of distribution:
(2.14) |
Lemma 2.5.
Let and be the solution of (2.14). Assume that is a nonnegative integer, then there exists depending only on and such that
(2.15) |
Proof.
We write (2.14)1 and (2.14)2 as
The symbol of the above system is
A direct calculation gives
This means that the boundary value problem (2.14) is elliptic in the sense of Agmon-Douglis-Nirenberg [Agmon:1964]. Moreover, the boundary condition is pure Dirichlet, and it is straightforward to verify that the boundary condition satisfies the complementing condition [Agmon:1964]. The regularity estimate (2.15) follows from [Agmon:1964]. ∎
A direct consequence of the above lemma is the following regularity estimate for problem (2.11).
Lemma 2.6.
Let be the solution of (2.11), there exists independent of such that
(2.16) |
Proof.
Using the standard elliptic regularity estimate, there exists a unique solution when . We introduce and , hence and satisfy (2.14) with and .
We turn to prove the regularity of problem (1.1) when . Let and be the solutions of (1.1) and (2.8), respectively. For any , integration by parts gives
(2.17) |
and
(2.18) |
where and . The boundary term in (2.18) is derived by the fact and
(2.19) | ||||
A combination of (2.17) and (2.18) leads to
which together with (2.9) yields
(2.20) |
This identity is the starting point of the proof.
We shall frequently use the following multiplicative trace inequality. There exists that depends only on such that
(2.21) |
Using Lemma 2.6, we condense the regularity of problem (1.1) to estimating and in various norms, with and .
Lemma 2.7.
There exists independent of and such that
(2.22) |
Proof.
The next lemma is crucial to prove Theorem 2.9, which transforms the estimate of in terms of besides a term concerning .
Lemma 2.8.
There exists independent of and such that
(2.23) |
Proof.
By [Danchin:2013]*Theorem 3.1 and Poincaré’s inequality, there exists such that , and
(2.24) |
Substituting into (2.20), multiplying the resulting identity by , we obtain
(2.25) |
By the regularity estimate (2.10), the first term may be bounded as
Using the triangle inequality and (2.10) again, we obtain
(2.26) |
Using (2.24) and the above inequality, we bound the second term as
Using (2.19) and the definition of , we rewrite the boundary term as
Recalling the trace inequality (2.21), we estimate the first boundary term as
Using (2.24), there exists independent of and such that
Using (2.26) to estimate and using (2.22) to bound , we obtain
A combination of the above three inequalities gives
We are ready to prove the regularity of problem (1.1).
Theorem 2.9.
There exists independent of and such that
(2.27) |
and
(2.28) |
The estimate (2.27) improves the known results [LiMingWang:2021]*Lemma 1 in two aspects. It clarifies the fact that the estimate is independent and it gives the convergence rate for the pressure. The rate is optimal even for the scalar counterpart; cf., [Tai:2001]*Lemma 5.1.
Proof.
Substituting into (2.20), we get
Using the regularity estimate (2.10), we bound the first term as
To bound the second term, we let in (2.19) and obtain
Invoking the trace inequality, using the fact and (2.10), we obtain
Substituting (2.23) into (2.22), we obtain,
Invoking (2.26) again, we get
A combination of the above three inequalities gives
A direct consequence of the above theorem is
Corollary 2.10.
There exists independent of and such that
(2.29) |
and
(2.30) |
3. A Family of Nonconforming Finite Elements
We introduce a family of finite elements to approximate the mixed variational problem (2.2). Let be a triangulation of with maximum mesh size . We assume all elements in are shape-regular in the sense of Ciarlet and Raviart [Ciarlet:1978]. We also assume that satisfies the inverse assumption: there exists such that for all . The space of piecewise vector fields is defined by
which is equipped with the broken norm
where For an interior edge shared by the triangles and , we define the jump of across as
where is the unit normal vector of towards the outside of . For , we set .
3.1. A family of finite elements
Our construction is motivated by the element proposed in [GuzmanLeykekhmanNeilan:2012]. Define the element with a triple by specifying as a triangle, and
(3.1) |
where and with the barycentric coordinates of .
Define
(3.2) |
and
(3.3) |
The degrees of freedom (DoFs) for are given by
(3.4) |
We plot the DoFs for in Figure 1.
Lemma 3.1.
The set is unisolvent.
Proof.
Firstly we show that if the DoFs (3.4) can determine an element in , then the element is unique. Suppose vanishes at the DoFs listed in (3.4), it suffices to show . Assume that
where , and , and . DoFs associated with are
The bubble space vanishes on this subset of DoFs by (3.2) and (3.3). The number of the DoFs is , which equals to the cardinality of . Hence .
A direct calculation gives
Taking in the above identity, we obtain on . Therefore, we write for certain . Using (3.2), we get
Taking in the above identity, we obtain . Therefore for certain . The last set of DoFs equals zero, i.e.,
Taking in the above identity, we obtain . So does .
We define a local interpolation operator as:
(3.5) |
Lemma 3.2.
There exists independent of such that for with , there holds
(3.6) |
where .
Proof.
Remark 3.3.
The element has a natural extension to three-dimensions by specifying as a tetrahedron, and
where is the element bubble function with the barycentric coordinates associated with the vertices for . is the face bubble function associated with the face .
3.2. Explicit representation for the bubble space
We clarify the structures of (3.2) and (3.3) associated with the set of DoFs (3.4)3 and the subset of (3.4)4 respectively, and derive the explicit formulations of the corresponding shape functions, which seems missing in the literature, while such explicit representations are useful for implementation. We firstly recall the following facts about the Jacobi polynomials [Szego:1975]. For any and nonnegative integers , there holds
(3.7) |
where
By [Szego:1975]*Eq. (4.3.3), we may write
(3.8) |
where is the Gamma function.
One of the explicit form for is
In particular,
Next we list certain facts about the Jacobi polynomials on the triangle [Dunkl:2014]*Section 2.4. For a triangle with vertices , any point is uniquely expressed as
Then is the barycentric coordinates of the point with respect to . For nonnegative integers such that , we define
(3.9) |
It is straightforward to verify . In particular,
and
For all , and nonnegative integers such that and , there holds
(3.10) | ||||
where is the standard reference triangle and
(3.11) | ||||
Using the notation , we may find that the expression (3.11) is equivalent to the one in [Dunkl:2014]*Eq. (2.4.3). The identity (3.10) illustrates that are mutually orthogonal bases of with respect to the weight .
Next we study the structure of the bubble spaces. For the barycentric coordinate function such that on , let and be the two other barycentric coordinates associated with the edges and , respectively. are chosen in a counterclockwise manner. The space can be clarified by the Jacobi polynomials with respect to the weight , while can be clarified by the Jacobi polynomials with respect to the weight , which are formulated in the following lemmas.
Lemma 3.4.
The space takes the form
Proof.
Motivated by the above lemma, we change the definition of DoFs for the bubble space from for any to
Lemma 3.5.
The shape functions for the bubble space associated with the above definition of DoFs are
with
(3.12) |
Proof.
A direct calculation gives
For . Using the relation (3.7), and noting that on , we obtain
This gives
Next we list the shape functions for the elements of low-order.
Example 3.6.
The bubble space for the lowest-order is
The shape functions associated with is
The bubble space for the case is
The shape functions associated with is
The shape functions associated with is
Lemma 3.7.
The space takes the forms
Proof.
Motivated by the above lemma, we change the definition of DoFs for from for any to
Lemma 3.8.
The shape functions for the bubble space associated with the above definition of DoFs are
with
Proof.
According to the definition of bubble space, we may have
Example 3.9.
The bubble space for the lowest-order case is span.
The shape functions associated with is .
The bubble space for the case is span.
The shape functions associated with is .
The shape functions associated with is .
Remark 3.10.
Based on the above results, we give the details for the element of the lowest-order, i.e., , which have been used in the numerical examples. The local finite element space
and DoFs
The shape functions associated with are
The shape functions associated with are
The shape functions associated with are
The shape functions associated with is .
4. The Mixed Finite Elements Approximation
In this part, we construct stable finite element pairs to approximate (2.2). We ignore the influence of the curved boundary for error estimates for brevity. Define
and . Let be the continuous Lagrangian finite element of order . We shall prove a uniform discrete B-B inequality for the pair .
The following rescaled trace inequality will be used later on: There exists independent of such that
(4.1) |
This inequality may be found in [BrennerScott:2008].
4.1. The mixed finite element approximation
We define the mixed finite element approximation problem as follows. Find and such that
(4.2) |
where
Note that , and this is a nonconforming method, we introduce the broken norm
Due to the continuity of , is a norm over .
The following broken Korn’s inequality was proved in [LiMingWang:2021]*Theorem 2:
which together with the first Korn’s inequality (2.6) gives
(4.3) |
It remains to prove the discrete B-B inequality for the pair . To this end, we construct a Fortin operator that is uniformly stable in the weighted norm [Winther:2013]. The key is to construct different Fortin operators for in different regimes.
Firstly we define an interpolation operator by , which satisfies
Lemma 4.1.
For all , there holds
(4.4) |
Proof.
Using the fact that , an integration by parts gives
(4.5) |
where we have used the identity (3.5)4 in the last step.
The operator is not Hbounded by (3.6), and we construct an Hbounded regularized interpolation operator as follows.
Lemma 4.2.
There exists an operator satisfying
(4.6) |
and if with , then
(4.7) |
The construction of is based on a regularized interpolation operator in [GuzmanLeykekhmanNeilan:2012] and the standard construction of the Fortin operator [Fortin:1977]. The operator is also well-defined for functions in .
Proof.
Define with and
where the regularized interpolation operator was constructed in [GuzmanLeykekhmanNeilan:2012]*Lemma 2, which satisfies
(4.8) |
The operator is defined for any element as
On each element , we have for any , and hence . A standard scaling argument gives
For any , using the fact that and , an integration by parts gives
where we have used the last property of . This gives (4.6).
We are ready to prove the following discrete B-B inequality.
Theorem 4.3.
There exists independent of and , such that
(4.9) |
Proof.
Using (2.3), for any , there exists such that
First, we consider the case with to be determined later on. By (4.7), we obtain
and
Combining the above inequality and using the inverse inequality for any , we obtain
Fix such that , we obtain
This gives
(4.10) |
The well-posedness of the mixed approximation problem (4.2) follows from the ellipticity of and the discrete B-B inequality of . We are ready to derive the error estimate.
4.2. Error estimates
To carry out the error estimate, we define the bilinear form as
for all and .
We prove the following inf-sup inequality for with the aid of the discrete B-B inequality (4.9).
Lemma 4.4.
Proof.
We are ready to prove error estimates.
Theorem 4.5.
There exists independent of and such that
(4.13) |
and
(4.14) |
Proof.
Let and with and , for any and ,
The boundedness of yields
Let be the interpolation of and be the order Lagrangian interpolation of , respectively. The standard interpolation error estimates in (3.6) gives
Note that
A standard estimate for the consistency error functional with trace inequality (4.1) gives
A combination of the above three inequalities, the discrete inf-sup condition (4.12) and the triangle inequalities gives (4.13).
Next, let and let be the Clément interpolation [Clement:1975] of , respectively. The interpolation error (4.7) and the error estimates for the Clément interpolation give
Corollary 4.6.
5. Numerical Examples
In this part, we report the numerical performance for the proposed element of the lowest-order, i.e., . We test the accuracy and robustness of the element pair for the nearly incompressible materials. All examples are carried out on the nonuniform mesh. We are interested in the case when the Poisson’s ratio is close to and we report the relative errors and the rates of convergence.
We let , and set Young’s modulus . The Lamé constants are determined by
We set for the ordinary cases, hence , and , and we set for the nearly incompressible materials, hence , and .
5.1. The first example
We test the performance of the element pair by solving a completely incompressible problem, which means . Let with
Therefore , and is independent of .
1/8 | 1/16 | 1/32 | 1/64 | |
1e+00 | 2.592e-01 | 1.333e-01 | 6.519e-02 | 3.246e-02 |
rate | 0.96 | 1.03 | 1.01 | |
1e-06 | 4.252e-02 | 1.159e-02 | 2.784e-03 | 6.918e-04 |
rate | 1.88 | 2.06 | 2.01 | |
1e+00 | 2.592e-01 | 1.333e-01 | 6.519e-02 | 3.246e-02 |
rate | 0.96 | 1.03 | 1.01 | |
1e-06 | 4.252e-02 | 1.159e-02 | 2.784e-03 | 6.918e-04 |
rate | 1.88 | 2.06 | 2.01 |
5.2. The second example
This example is motivated by [Wihler:2006], which admits a singular solution. The exact solution expressed in the polar coordinates as
where
and ,
It may be verified that for a small number . A direct calculation gives that , while it is nearly incompressible because
1/8 | 1/16 | 1/32 | 1/64 | |
1e+00 | 1.062e-01 | 7.554e-02 | 5.355e-02 | 3.792e-02 |
rate | 0.49 | 0.50 | 0.50 | |
1e-06 | 2.809e-03 | 1.001e-03 | 3.549e-04 | 1.257e-04 |
rate | 1.49 | 1.50 | 1.50 | |
1e+00 | 1.149e-01 | 8.200e-02 | 5.824e-02 | 4.135e-02 |
rate | 0.49 | 0.49 | 0.49 | |
1e-06 | 4.399e-03 | 1.567e-03 | 5.558e-04 | 1.968e-04 |
rate | 1.49 | 1.50 | 1.50 |
It follows from Table 3 that the rates of convergence are sub-optimal. It is reasonable because the solution is singular, which is similar to the results in [Wihler:2006]. The element pair is robust for the nearly incompressible materials.
5.3. The third example
In the last example, we test a problem with strong boundary layer effects. Such effects have been frequently observed in the stain elasticity model [Engel:2002, LiMingShi:2017, LiaoM:2019, LiMingWang:2021]. It is shown that the numerical solution converges to the solution of (2.8) when .
When , the boundary value problem (1.1) reduces to (2.8). Let with
be the solution of problem (2.8). The source term is computed from (2.8). A direct calculation gives that , and is independent of . The exact solution for (1.1) is unknown, while it has strong boundary layer effects. In this case, we take , and report the relative error .
1/8 | 1/16 | 1/32 | 1/64 | |
1e-04 | 1.311e-01 | 8.966e-02 | 6.299e-02 | 4.476e-02 |
rate | 0.55 | 0.51 | 0.49 | |
1e-06 | 1.311e-01 | 8.960e-02 | 6.283e-02 | 4.432e-02 |
rate | 0.55 | 0.51 | 0.50 | |
1e-04 | 1.312e-01 | 8.968e-02 | 6.300e-02 | 4.476e-02 |
rate | 0.55 | 0.51 | 0.49 | |
1e-06 | 1.312e-01 | 8.963e-02 | 6.284e-02 | 4.432e-02 |
rate | 0.55 | 0.51 | 0.50 |
It follows from Table 4 that the rate of convergence for the element pair changes to because of the boundary layer effects, which is consistent with the theoretical result. The element is still robust when the solution has strong boundary layer effects in the nearly incompressible limit.