Skip to main content
Erschienen in: Computational Mechanics 6/2022

Open Access 11.08.2022 | Original Paper

Lagrange and \(H({\text {curl}},{{\mathcal {B}}})\) based finite element formulations for the relaxed micromorphic model

verfasst von: Jörg Schröder, Mohammad Sarhil, Lisa Scheunemann, Patrizio Neff

Erschienen in: Computational Mechanics | Ausgabe 6/2022

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Modeling the unusual mechanical properties of metamaterials is a challenging topic for the mechanics community and enriched continuum theories are promising computational tools for such materials. The so-called relaxed micromorphic model has shown many advantages in this field. In this contribution, we present significant aspects related to the relaxed micromorphic model realization with the finite element method (FEM). The variational problem is derived and different FEM-formulations for the two-dimensional case are presented. These are a nodal standard formulation \(H^1({{\mathcal {B}}}) \times H^1({{\mathcal {B}}})\) and a nodal-edge formulation \(H^1({{\mathcal {B}}}) \times H({\text {curl}}, {{\mathcal {B}}})\), where the latter employs the Nédélec space. In this framework, the implementation of higher-order Nédélec elements is not trivial and requires some technicalities which are demonstrated. We discuss the computational convergence behavior of Lagrange-type and tangential-conforming finite element discretizations. Moreover, we analyze the characteristic length effect on the different components of the model and reveal how the size-effect property is captured via this characteristic length parameter.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Metamaterials are receiving tremendous attention in both academia and industry due to their unconventional mechanical properties. These are not solely governed by the bulk mechanical properties but also by the geometry of the unit cells which can be designed to attain the desired functionality, see [14, 19, 53, 55, 56]. Moreover, recent advances of additive manufacturing (AM, or 3D printing) techniques are empowering the fabrication process of three-dimensional architected metamaterials, c.f. [16, 20, 32, 42]. To simplify the design process of novel metamaterials, suitable computational tools are needed to capture their unprecedented effective mechanical properties. The classical Cauchy-Boltzmann theory and first-order homogenization procedures often fail to describe the mechanical macroscopic behavior of mechanical metamaterials since they exhibit the size-effect phenomenon, i.e. small specimens are stiffer than big specimens, and therefore other generalized theories are needed such as the classical Mindlin-Eringen micromorphic theory [11, 12, 21, 22, 29, 52], the Cosserat theory [7, 35, 36], gradient elasticity [2, 13, 30] or others.
The relaxed micromorphic model, which we adopt in this work, has been introduced recently in [15, 37, 38]. It keeps the full kinematics of the micromorphic theory but employs the matrix Curl operator of a non-symmetric second-order micro-distortion field for the curvature measurement. The relaxed micromorphic model reduces the complexity of the classical micromorphic theory by decreasing the number of material parameters and has shown many advantages such as the separation of material parameters into scale-dependent and scale-independent ones, see for example [9]. Furthermore, it has already been used to obtain the main mechanical characteristics (stiffness, anisotropy, dispersion) of targeted metamaterials for many well-posed dynamical and statical problems, e.g. [1, 3, 4, 2328, 39]. Recently, the scale-independent short range elastic parameters in the relaxed micromorphic model were determined for artificial periodic microstructures in [40] which are used to capture band-gaps as a dynamical property of mechanical metamaterial in [9]. Analytical solutions of the relaxed micromorphic compared to solutions of other generalized continua for some essential boundary value problems, i.e. pure shear, bending, torsion and uniaxial tension, are discussed in [4447], emphasizing the validity of the relaxed micromorphic model for small sizes (bounded stiffness), where most of the other generalized continua exhibit unphysical stiffness properties.
As a result of employing the matrix Curl operator of the micro-distortion field for the curvature measurement, the relaxed micromorphic model seeks the solution of the micro-distortion in \(H({\text {curl}},{{\mathcal {B}}})\), while the displacement solution is still in \(H^1({{\mathcal {B}}})\). The appropriate finite elements of such case must be conforming in \(H({\text {curl}},{{\mathcal {B}}})\) (tangentially conforming). The first formulation of edge elements was presented in [43]. In fact, the name “edge” elements was used because the degrees of freedom (dofs) are associated only with edges for a first-order approximation. \(H({\text {curl}},{{\mathcal {B}}})\)-conforming finite elements of first kind were introduced in [33] and second kind in [34], which are comparable with \(H({\text {div}},{{\mathcal {B}}})\)-conforming elements of first kind in [43] and second kind in [6]. An extension to elements with curved edges, based on covariant projections, was developed by [8]. A general implementation of Nédélec elements of first kind is presented in [41] and a detailed review about \(H({\text {div}},{{\mathcal {B}}})\)- and \(H({\text {curl}},{{\mathcal {B}}})\)-conforming finite elements is available in [17] and [48]. Furthermore, hierarchical \(H({\text {curl}})\)-conforming finite elements are used to solve Maxwell boundary and eigenvalue problems in [49]. A \(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) finite element formulation for a simplified anti-plane shear case of the relaxed micromorphic model utilizing a scalar displacement field and a vectorial micro-distortion field is available in [50].
In our work, we demonstrate the main technologies related to the finite element realization of the theoretically-sound relaxed micromorphic model. The proper finite element approximation of the micro-distortion field is the Nédélec space which utilizes tangential-conforming vectorial shape functions. We provide a comprehensive description of the construction of \(H^1({{\mathcal {B}}}) \times H({\text {curl}}, {{\mathcal {B}}})\) elements with Nédélec formulation of first kind on triangular and quadrilateral meshes. Six finite elements are built, which differ in the approximation space of the micro-distortion: two triangular elements with first- and second-order Nédélec formulation, two quadrilateral elements with first- and second-order Nédélec formulation, and two nodal triangular elements with standard first- and second-order Lagrangian formulation. The paper is organized as follows. In Sect. 2, we introduce the relaxed micromorphic model and derive the variational problem with the resulting strong forms and associated boundary conditions which are modulated in a physical point of view by the so-called consistent coupling condition. We cover in Sect. 3 the main components of the implementation of standard nodal and nodal-edge elements. Two numerical examples are introduced in Sect. 4. The first numerical example is designed to check the convergence behavior of the different finite elements when the solution is discontinuous in the micro-distortion field. We investigate the influence of the characteristic length in a second example which covers the size-effect property. We conclude the paper in Sect. 5.

2 The relaxed micromorphic model

The relaxed micromorphic model is a continuum model which describes the kinematics of a material point using a displacement vector \({{\varvec{u}}}:{{\mathcal {B}}}\subseteq {\mathbb {R}}^3\rightarrow {\mathbb {R}}^3\) and a non-symmetric micro-distortion field \({{\varvec{P}}}:{{\mathcal {B}}}\subseteq {\mathbb {R}}^3\rightarrow {\mathbb {R}}^{3\times 3}\). Both are defined for the static case by minimizing the potential
$$\begin{aligned} \Pi ({{\varvec{u}}},{{\varvec{P}}})= & {} \mathop {\int }\limits _{{\mathcal {B}}}W\left( \nabla {{\varvec{u}}},{{\varvec{P}}},{\text {Curl}}{{\varvec{P}}}\right) \, - {\overline{{{\varvec{f}}}}}\cdot {{{\varvec{u}}}} \, - {\overline{{{\varvec{M}}}}} : {{\varvec{P}}}\,{\text {d}}V \nonumber \\&- \mathop {\int }\limits _{\partial {{\mathcal {B}}}_t} {\overline{{{\varvec{t}}}}} \cdot {{\varvec{u}}}\,{\text {d}}A \longrightarrow \ \min \,, \end{aligned}$$
(1)
with \(({{\varvec{u}}},{{\varvec{P}}})\in H^1({{\mathcal {B}}})\times H({\text {curl}},{{\mathcal {B}}})\). The vector \({{\overline{{{\varvec{f}}}}}}\) and tensor \({\overline{{{\varvec{M}}}}}\) describe, respectively, the given body force and body moment, while \({{\overline{{{\varvec{t}}}}}}\) is the traction vector acting on the boundary \(\partial {{\mathcal {B}}}_t \subset \partial {{\mathcal {B}}}\). The elastic energy density W reads
$$\begin{aligned} \begin{aligned}&W\left( \nabla {{\varvec{u}}},{{\varvec{P}}},{\text {Curl}}{{\varvec{P}}}\right) \\&= \, \dfrac{1}{2} ( {\text {sym}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}] : {\mathbb {C}}_{{\mathrm {e}}}: {\text {sym}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}] \\&\quad + {\text {sym}}{{\varvec{P}}}: {\mathbb {C}}_{{\mathrm {micro}}}: {\text {sym}}{{\varvec{P}}}\, \\&\quad + {\text {skew}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}] : {\mathbb {C}}_{{\mathrm {c}}}: {\text {skew}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}] \\&\quad + \mu \, L_{{\mathrm {c}}}^2 \, {\text {Curl}} {{\varvec{P}}}: {\mathbb {L}}:{\text {Curl}} {{\varvec{P}}}) \,. \\ \end{aligned} \end{aligned}$$
(2)
Here, \({\mathbb {C}}_{{\mathrm {micro}}},{\mathbb {C}}_{{\mathrm {e}}}\) are fourth-order positive definite standard elasticity tensors, \({\mathbb {C}}_{{\mathrm {c}}}\) is a fourth-order positive semi-definite rotational coupling tensor, \({\mathbb {L}}\) is a fourth-order tensor, \(L_{{\mathrm {c}}}\) is a non-negative parameter describing the characteristic length scale and \(\mu \) is a typical shear modulus. The characteristic length parameter plays a significant role in the relaxed micromorphic model. This parameter is related to the size of the microstructure and determines its influence on the macroscopic mechanical behavior. The characteristic length allows to scale the number of considered unit cells keeping all remaining parameters of the model scale-independent. It accounts for non-localities in the considered metamaterial where the deformation of each unit cell is influenced by deformations and motions of the neighbouring cells. A relation of the relaxed micromorphic model to the classical Cauchy model was shown in [40] for limiting values of \(L_{{\mathrm {c}}}\), which we can also observe in our numerical examples.
The variation of the potential with respect to the displacement field, i.e. \(\delta _{{{\varvec{u}}}} \Pi = 0\), with
$$\begin{aligned} \delta _{{{\varvec{u}}}} \Pi= & {} \mathop {\int }\limits _{{{\mathcal {B}}}} \{ \underbrace{{\mathbb {C}}_{{\mathrm {e}}}: {\text {sym}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}]+ {\mathbb {C}}_{{\mathrm {c}}}: {\text {skew}}[{ \nabla {{\varvec{u}}}- {{\varvec{P}}}}]}_{ \textstyle =:{\varvec{\sigma }}} \} : {\nabla \delta {{\varvec{u}}}} \nonumber \\&- {{\overline{{{\varvec{f}}}}}} \cdot \delta {{\varvec{u}}}\, {\text {d}}V - \mathop {\int }\limits _{\partial {{\mathcal {B}}}_t} {{\overline{{{\varvec{t}}}}}} \cdot \delta {{\varvec{u}}}\, {\text {d}}A\,, \end{aligned}$$
(3)
leads, after integration by parts and applying the divergence theorem, to the weak form
$$\begin{aligned} \delta _{{{\varvec{u}}}} \Pi = \mathop {\int }\limits _{{{\mathcal {B}}}} \{ {\text {div}}{\varvec{\sigma }}+ {{\overline{{{\varvec{f}}}}}} \} \cdot \delta {{\varvec{u}}}\, {\text {d}}V = 0 \, , \end{aligned}$$
(4)
where \({\varvec{\sigma }}\) denotes the non-symmetric force stress tensor. The associated strong form with related boundary conditions reads
$$\begin{aligned}&{\text {div}}{\varvec{\sigma }}+{{\overline{{{\varvec{f}}}}}} = {\mathbf{0}}\, \quad {\text {with}} \quad {{\varvec{u}}}= {\overline{{{\varvec{u}}}}} \quad {\text {on}} \quad \partial {{\mathcal {B}}}_u\, \quad {\text {and}}\nonumber \\&\quad {{\overline{{{\varvec{t}}}}}} = {\varvec{\sigma }}\cdot {{\varvec{n}}}\quad {\text {on}} \quad \partial {{\mathcal {B}}}_t \,, \end{aligned}$$
(5)
satisfying \(\partial {{\mathcal {B}}}_u \cap \partial {{\mathcal {B}}}_t = \emptyset \) and \(\partial {{\mathcal {B}}}_u \cup \partial {{\mathcal {B}}}_t = \partial {{\mathcal {B}}}\) and \({{\varvec{n}}}\) is the outward normal on the boundary. In a similar way, the variation of the potential with respect to the micro-distortion field, i.e. \(\delta _{{{\varvec{P}}}} \Pi = 0\), with
$$\begin{aligned} \delta _{{{\varvec{P}}}} \Pi= & {} \mathop {\int }\limits _{{{\mathcal {B}}}} \{ {\varvec{\sigma }}- \underbrace{{\mathbb {C}}_{{\mathrm {micro}}}: {\text {sym}}{{\varvec{P}}}}_{\textstyle =:{\varvec{\sigma }}_{\text {micro}} } + {{\overline{{{\varvec{M}}}}}}\} : \delta {{{\varvec{P}}}}\nonumber \\&- \underbrace{\mu \, L_{{\mathrm {c}}}^2 ({\mathbb {L}}: {\text {Curl}}{{\varvec{P}}})}_{\textstyle =:{{\varvec{m}}}} : {\text {Curl}}\delta {{{\varvec{P}}}} \, {\text {d}} V\,, \end{aligned}$$
(6)
yields after integration by parts and applying Stokes’ theorem
$$\begin{aligned} \begin{aligned} \delta _{{{\varvec{P}}}} \Pi&= \mathop {\int }\limits _{{{\mathcal {B}}}} \{ {\varvec{\sigma }}- {\varvec{\sigma }}_{\text {micro}} - {\text {Curl}}{{\varvec{m}}}+ {{\overline{{{\varvec{M}}}}}}\} : \delta {{{\varvec{P}}}} \, {\text {d}} V \\&\quad + \mathop {\int }\limits _{ \partial {{\mathcal {B}}}} \left\{ \sum _{i=1}^3 \left( {{\varvec{m}}}^i \times \delta {{{\varvec{P}}}}^i \right) \cdot {{\varvec{n}}}\right\} \;\; {\text {d}}A = 0\,, \end{aligned} \end{aligned}$$
(7)
where \({\varvec{\sigma }}_{\text {micro}} \) and \({{\varvec{m}}}\) are the micro- and moment stresses, respectively, and \({{\varvec{m}}}^i\) and \(\delta {{\varvec{P}}}^i\) denote row vectors of the associated second-order tensors. Using the identity of the scalar triple vector product
$$\begin{aligned} ({{\varvec{a}}}\times {{\varvec{b}}}) \cdot {{\varvec{c}}}= ({{\varvec{c}}}\times {{\varvec{a}}}) \cdot {{\varvec{b}}}= ({{\varvec{b}}}\times {{\varvec{c}}}) \cdot {{\varvec{a}}}\,, \end{aligned}$$
(8)
allows for the reformulation
$$\begin{aligned}&\mathop {\int }\limits _{ \partial {{\mathcal {B}}}} \left\{ \sum _{i=1}^3 \left( {{\varvec{m}}}^i \times \delta {{{\varvec{P}}}}^i \right) \cdot {{\varvec{n}}}\right\} \;\; {\text {d}}A\nonumber \\&= \mathop {\int }\limits _{ \partial {{\mathcal {B}}}_P} \left\{ \sum _{i=1}^3 \left( \delta {{{\varvec{P}}}}^i \times {{\varvec{n}}}\right) \cdot {{\varvec{m}}}^i \right\} \;\; {\text {d}}A \nonumber \\&- \mathop {\int }\limits _{ \partial {{\mathcal {B}}}_m} \left\{ \sum _{i=1}^3 \left( {{\varvec{m}}}^i \times {{\varvec{n}}}\right) \cdot \delta {{{\varvec{P}}}}^i \right\} \;\; {\text {d}}A \,. \end{aligned}$$
(9)
The associated strong form reads
$$\begin{aligned} {\text {Curl}}{{\varvec{m}}}= {\varvec{\sigma }}- {\varvec{\sigma }}_{\text {micro}} + {{\overline{{{\varvec{M}}}}}} \,, \end{aligned}$$
(10)
with related boundary conditions, formulated in terms of the row vectors \({{\varvec{P}}}^i\) and \({{\varvec{m}}}^i\) of the associated second-order tensors,
$$\begin{aligned}&\sum _{i=1}^3 {{{\varvec{P}}}}^i \times {{\varvec{n}}}= {\overline{{{\varvec{t}}}}}_p \quad {\text {on}} \quad \partial {{\mathcal {B}}}_P\quad {\text {and by definition}} \nonumber \\&\sum _{i=1}^3 {{\varvec{m}}}^i \times {{\varvec{n}}}= {\mathbf{0}}\quad {\text {on}} \quad \partial {{\mathcal {B}}}_m\,, \end{aligned}$$
(11)
where \(\partial {{\mathcal {B}}}_P\cap \partial {{\mathcal {B}}}_m = \emptyset \) and \(\partial {{\mathcal {B}}}_P\cup \partial {{\mathcal {B}}}_m= \partial {{\mathcal {B}}}\). A dependency between the displacement field and the micro-distortion field on the boundary was proposed by [40] and subsequently considered in [10, 45, 46, 50]. This so-called consistent coupling condition is defined by
$$\begin{aligned}&{{\varvec{P}}}\cdot {\varvec{\tau }}= \nabla {{\varvec{u}}}\cdot {\varvec{\tau }}\, \Leftrightarrow \, {{{\varvec{P}}}}^i \times {{\varvec{n}}}= {\nabla {{\varvec{u}}}}^i \times {{\varvec{n}}}\nonumber \\&\quad {\text {for}} \quad i=1,2,3 \quad {\text {on}}\quad \partial {{\mathcal {B}}}_P= \partial {{\mathcal {B}}}_u \,, \end{aligned}$$
(12)
where \({\varvec{\tau }}\) is a tangential vector on the Dirichlet boundary and \({\nabla {{\varvec{u}}}}^i\) are the row vectors of \(\nabla {{\varvec{u}}}\). This condition relates the projection of the displacement gradient on the tangential plane of the boundary to the respective parts of the micro-distortion.
The first strong form in Equation (5) represents a generalized balance of linear momentum (force balance) while the second strong form in Equation (10) outlines a generalized balance of angular momentum (moment balance). The generalized moment balance invokes the Cosserat theory with the \({\text {Curl}}{\text {Curl}}\) operator rising from the matrix \({\text {Curl}}\) operator of the second-order moment stress \({{\varvec{m}}}\). In comparison to the classical micromorphic model, see [11, 37], the relaxed micromorphic model uses the same kinematical measures but employs a curvature measure from the Cosserat theory, see [36]. The micro-distortion field has the following general form for the three-dimensional case
$$\begin{aligned} {{\varvec{P}}}= & {} \left[ \begin{array}{c} ({{\varvec{P}}}^{1})^T \\ ({{\varvec{P}}}^{2})^T \\ ({{\varvec{P}}}^{3})^T \\ \end{array}\right] = \left[ \begin{array}{c c c} P_{11} &{} P_{12} &{} P_{13} \\ P_{21} &{} P_{22} &{} P_{23} \\ P_{31} &{} P_{32} &{} P_{33} \\ \end{array}\right] \nonumber \\&\quad {\text {with}} \quad {{\varvec{P}}}^{i} = \left[ \begin{array}{c} P_{i1} \\ P_{i2} \\ P_{i3} \end{array} \right] , \quad {i=1,2,3 \,,}\nonumber \\ \end{aligned}$$
(13)
where \({{\varvec{P}}}^i\) denotes the row vectors of \({{\varvec{P}}}\). We let the Curl operator act on the row vectors of the micro-distortion field \({{\varvec{P}}}\), i.e.,
$$\begin{aligned} {\text {Curl}}{{\varvec{P}}}= & {} \left[ \begin{array}{c} ({\text {curl}}{{\varvec{P}}}^{1})^T \\ ({\text {curl}}{{\varvec{P}}}^{2})^T \\ ({\text {curl}}{{\varvec{P}}}^{3})^T \\ \end{array}\right] \nonumber \\= & {} \left[ \begin{array}{c|c|c} P_{13,2} - P_{12,3} &{} P_{11,3} - P_{13,1} &{} P_{12,1} - P_{11,2} \\ P_{23,2} - P_{22,3} &{} P_{21,3} - P_{23,1} &{} P_{22,1} - P_{21,2} \\ P_{33,2} - P_{32,3} &{} P_{31,3} - P_{33,1} &{} P_{32,1} - P_{31,2} \end{array}\right] .\nonumber \\ \end{aligned}$$
(14)
For the two-dimensional case, the micro-distortion field and its Curl operator are reduced to
$$\begin{aligned} {{\varvec{P}}}= & {} \left[ \begin{array}{c} ({{\varvec{P}}}^{1})^T \\ ({{\varvec{P}}}^{2})^T \\ {\mathbf{0}}^T \end{array}\right] = \left[ \begin{array}{c c c } P_{11} &{} P_{12} &{} 0\\ P_{21} &{} P_{22} &{} 0\\ 0 &{} 0 &{} 0 \end{array}\right] \quad {\text {and}} \nonumber \\ {\text {Curl}}{{\varvec{P}}}= & {} \left[ \begin{array}{c|c|c} 0 &{} 0 &{} P_{12,1} - P_{11,2} \\ 0 &{} 0 &{} P_{22,1} - P_{21,2} \\ 0 &{} 0 &{} 0 \end{array}\right] . \end{aligned}$$
(15)

3 Approximation spaces

3.1 Nodal elements \(({{\varvec{u}}},{{\varvec{P}}}) \in H^1({{\mathcal {B}}}) \times H^1({{\mathcal {B}}})\)

We introduce the formulation of a standard nodal element utilizing Lagrange-type shape functions for both displacement and micro-distortion field, see for example [54]. Let us assume that there are \(n^u\) nodes in each element for the discretization of the displacement field \({{\varvec{u}}}\) and \(n^P\) nodes for micro-distortion field \({{\varvec{P}}}\) in two dimensions. Geometry and displacement field are approximated employing related Lagrangian shape functions \(N^u_I\) defined in the parameter space with natural coordinates \({\varvec{\xi }}= \{ \xi , \eta \}\) by
$$\begin{aligned} {{\varvec{X}}}_h = \sum _{I=1}^{n^u} N^u_I({\varvec{\xi }})\, {{\varvec{X}}}_I\,, \qquad {{\varvec{u}}}_h = \sum _{I=1}^{n^u} N^u_I({\varvec{\xi }}) \, {{\varvec{d}}}^u_I\,, \end{aligned}$$
(16)
where \({{\varvec{X}}}_I\) are the coordinates of displacement node I and \({{\varvec{d}}}^u_I\) are its displacement degrees of freedom. The deformation gradient is obtained in physical space by
$$\begin{aligned}&\nabla {{\varvec{u}}}_h = \sum _{I=1}^{n^u} {{\varvec{d}}}^u_I \otimes \nabla N^u_I({\varvec{\xi }})\, \quad {\text {with}}\nonumber \\&\nabla {N^u_I({\varvec{\xi }})} = {{\varvec{J}}}^{-T} \cdot \nabla _{\varvec{\xi }}N^u_{I}\,, \end{aligned}$$
(17)
where \({{\varvec{J}}}= \frac{\partial {{\varvec{X}}}}{\partial {\varvec{\xi }}}\) is the Jacobian, \(\nabla \) and \(\nabla _{\varvec{\xi }}\) denote gradient operators with respect to \({{\varvec{X}}}\) and \({\varvec{\xi }}\), respectively. The micro-distortion field \({{\varvec{P}}}\) for the 2D case is approximated using the relevant scalar shape functions \(N^P_I\)
$$\begin{aligned}&{{\varvec{P}}}_h^1 = \left[ \begin{array}{c} P_{11} \\ P_{12} \\ \end{array} \right] = \sum _{I=1}^{n^P} N^P_I({\varvec{\xi }})\, {{\varvec{d}}}^{P^1}_I\,,\nonumber \\&{{\varvec{P}}}_h^2 = \left[ \begin{array}{c} P_{21} \\ P_{22} \\ \end{array} \right] = \sum _{I=1}^{n^P} N^P_I({\varvec{\xi }}) \,{{\varvec{d}}}^{P^2}_I\,, \end{aligned}$$
(18)
where \({{\varvec{d}}}^{P^1}_I\) and \({{\varvec{d}}}^{P^2}_I\) are the micro-distortion row vectors degrees of freedom of node I. In order to calculate the Curl of \({{\varvec{P}}}\), the gradient of the row vectors in physical space can be calculated by
$$\begin{aligned} \nabla {{\varvec{P}}}_h^{i} = {{\varvec{J}}}^{-T} \cdot \nabla _{\varvec{\xi }}{{\varvec{P}}}_h^{i} \quad {\text {for}} \quad i=1,2 \end{aligned}$$
(19)
and the rotation of the vector \({{\varvec{P}}}_h^{i}\) reads
$$\begin{aligned} {\text {curl}}^{2D}{{{\varvec{P}}}_h^{i}} = (P_h)_{i2,1} - (P_h)_{i1,2} \quad {\text {for}} \quad i=1,2 \,. \end{aligned}$$
(20)

3.2 Nodal-edge elements \(({{\varvec{u}}},{{\varvec{P}}}) \in H^1({{\mathcal {B}}}) \times H({\text {curl}}, {{\mathcal {B}}})\)

The here presented formulation uses different spaces to describe the micro-distortion field. The geometry and displacement field are approximated in standard Lagrange space as in Equation (16). For the micro-distortion field, its solution is in \( H({\text {curl}},{{\mathcal {B}}})\) and the suitable finite element space is known as Nédélec space, see [33, 34]. In this work, we choose the Nédélec space of first kind. For more details the reader is referred to [5, 17, 31, 48]. Nédélec formulations use vectorial shape functions which satisfy tangential continuity at element interfaces. Lowest-order two-dimensional Nédélec elements are depicted in Fig. 1.
Triangular Nédélec elements of order k are based on the space
$$\begin{aligned}&\left[ \mathcal{ND}\mathcal{}^\triangle \right] ^{2}_k = ({\mathbb {P}}_{k-1})^2 \oplus {{\varvec{S}}}_k \quad {\text {with}}\nonumber \\&{{\varvec{S}}}_k = \{ {{\varvec{p}}}\in ({\tilde{{\mathbb {P}}}}_k)^2 \,|\, {{\varvec{p}}}\cdot {\varvec{\xi }}= 0 \}\,, \end{aligned}$$
(21)
where \({\mathbb {P}}_{k-1}\) is the linear space of polynomials of degree \(k-1\) or less and \({\tilde{{\mathbb {P}}}}_{k}\) is the linear space of homogeneous polynomials of degree k. Equivalently, the space can be characterized by
$$\begin{aligned} \left[ \mathcal{ND}\mathcal{}^\triangle \right] ^{2}_k = ({\mathbb {P}}_{k-1})^2 \oplus {\tilde{{\mathbb {P}}}}_{k-1} \left[ \begin{array}{c} - \eta \\ \xi \end{array} \right] . \end{aligned}$$
(22)
The dimension of this linear space is \( k(k+2) \). Quadrilateral Nédélec elements of order k are based on the linear space
$$\begin{aligned}&\left[ \mathcal{ND}\mathcal{}^\square \right] ^{2}_k = \left[ \begin{array}{c} Q_{k-1,k} \\ Q_{k,k-1} \end{array} \right] \quad {\text {where}} \nonumber \\&\quad Q_{m,n} = {\text {span}}\{ \xi ^i \eta ^j \,|\, i \le m, j \le n \}, \end{aligned}$$
(23)
with \( \dim \left( \left[ \mathcal{ND}\mathcal{}^\square \right] ^{2}_k \right) = 2k(k+1) \). The vectorial shape functions \({{\varvec{v}}}^k\) in parametric space are obtained by constructing a linear system of equations based on a set of inner and outer dofs. For the 2D case, the outer dofs of an edge \(e_i\) are defined by the integral
$$\begin{aligned} m^{e_i}_{j} ({{\varvec{v}}}^k ) = \mathop {\int }\limits _{e_i} ({{\varvec{v}}}^k \cdot {{\varvec{t}}}_i) \, r_{j} \, {\text {d}} s \,, \quad \forall \, r_j \in {\mathbb {P}}_{k-1}(e_i)\, \end{aligned}$$
(24)
where \(r_{j}\) is a polynomial \({\mathbb {P}}_{k-1}\) along edge \(e_i\) and \({{\varvec{t}}}_i\) is the normalized tangential vector of edge \(e_i\). The inner dofs are introduced for triangular elements by
$$\begin{aligned} m^{\text {inner}}_{i} ({{\varvec{v}}}^k ) = \mathop {\int }\limits _{{{\mathcal {B}}}_e} {{\varvec{v}}}^k \cdot {{\varvec{q}}}_{i} \, {\text {d}} a \,, \quad \forall \, {{\varvec{q}}}_i \in ({\mathbb {P}}_{k-2} ({{\mathcal {B}}}_e))^2\, , \end{aligned}$$
(25)
while they are given for quadrilateral elements by
$$\begin{aligned} m^{\text {inner}}_{i} ({{\varvec{v}}}^k ) = \mathop {\int }\limits _{{{\mathcal {B}}}_e} {{\varvec{v}}}^k \cdot {{\varvec{q}}}_{i} \, {\text {d}} a \,, \quad \forall \, {{\varvec{q}}}_i \, \in \left[ \begin{array}{c} Q_{k-1,k-2} ({{\mathcal {B}}}_e) \\ Q_{k-2,k-1} ({{\mathcal {B}}}_e) \end{array} \right] .\nonumber \\ \end{aligned}$$
(26)
The scalar-valued and vectorial functions \(r_j\) and \({{\varvec{q}}}_i\) are linearly independent polynomials which are chosen as Lagrange polynomials in our work. For lowest-order element (\(k=1\)), only outer dofs occur. For higher-order elements (\(k \ge 2\)), the number of outer dofs is increased and additional inner dofs are introduced. E.g. for the \(\left[ \mathcal{ND}\mathcal{}^\triangle \right] ^{2}_2\) with a dimension 8, we have 6 outer dofs and 2 inner ones. The derivations of \(H({\text {curl}},{{\mathcal {B}}})\)-conforming vectorial shape functions are shown in Appendix B. Mapping vectorial shape functions \( {{\varvec{v}}}^k_I\) from the parametric space to \({\hat{{\varvec{\psi }}}}^k_I \) in the physical space must conserve the tangential continuity property. This is guaranteed by using the covariant Piola transformation, see for example [48], which reads
$$\begin{aligned} {\hat{{\varvec{\psi }}}}^k_I = \, {{\varvec{J}}}^{-T} \cdot {{\varvec{v}}}^k_I\, \quad {\text {and}} \quad {\text {curl}}{{\hat{{\varvec{\psi }}}}^k_I} = \frac{1 }{{\text {det}}{{{\varvec{J}}}}} {{\varvec{J}}}\cdot {\text {curl}}_{{\varvec{\xi }}}{{\varvec{v}}}^k_I\,. \end{aligned}$$
(27)
For our implementation of \(H({\text {curl}},{{\mathcal {B}}})\)-conforming elements, we modify the mapping to enforce the required orientation of the degrees of freedom at inter-element boundaries and to attach a direct physical interpretation to the Neumann-type boundary conditions. Hence, two additional parameters, \(\alpha \) and \(\beta \), appear for the vectorial shape functions associated with edge dofs
$$\begin{aligned} {{{\varvec{\psi }}}}^k_I = \alpha _I \beta _I {{\hat{{\varvec{\psi }}}}}^k_I \quad {\text {and}} \quad {\text {curl}}{{{\varvec{\psi }}}^k_I} = \alpha _I \beta _I {\text {curl}}{{{\hat{{\varvec{\psi }}}}}^k_I} \,, \end{aligned}$$
(28)
where \( \alpha _I = \pm 1\) is the orientation consistency function which ensures that on an edge, belonging to two neighboring finite elements, a positive tangential flux direction is defined. Therefore, this positive tangential direction is defined based on a positive x-coordinate. A tangential component pointing in negative x-direction is multiplied by a value \(\alpha _I = -1\) to obtain an overall positive tangential flux on each edge. If the tangential direction has no projection on the x-axis, the same procedure is employed in y-direction. Fig. 2 illustrates an example of calculating the orientation parameter values of two neighboring elements.
The normalization parameter \(\beta _I\) enforces that the sum of vectorial shape functions \({{\varvec{\psi }}}^k_I \) at a common edge scalar multiplied with the associated tangential vector has to be equal one in physical space. Furthermore, the sum of the shape functions belonging to one edge scalar multiplied with the tangential vector of the other edges must vanish. These conditions are reflected by
$$\begin{aligned}&{\varvec{\tau }}_I \cdot \sum _J {{{\varvec{\psi }}}}_J^k \bigg |_{E_I} \equiv 1 \quad \, {\text {if}} \quad I=J \quad {\text {and}} \, \nonumber \\&{\varvec{\tau }}_I \cdot \sum _J {{{\varvec{\psi }}}}_J^k \bigg |_{E_I} \equiv 0 \quad {\text {if}} \quad I \ne J \,. \end{aligned}$$
(29)
Here, \(\sum _J {{{\varvec{\psi }}}}_J^k \bigg |_{E_I} \) is the sum of shape vectors related to outer dofs of an edge \(E_J\) evaluated on edge \(E_I\) and \({\varvec{\tau }}_I\) is the normalized tangential vector of an edge \(E_I\) where E denotes the edges in the physical space. Based on Equations (28)\(_1\) and (29)\(_1\), we compute straightforward the parameters \(\beta _I\). In detail, we get for first- and second-order elements
$$\begin{aligned} \beta _I = L_I \quad {\text {and}} \quad \beta _I = \frac{L_I}{2}\,, \end{aligned}$$
(30)
respectively, where \(L_I\) denotes the length of edge \(E_I\) in physical space. For the 2D case, the rotation of vectorial shape functions only has one active component out of the plane which reads
$$\begin{aligned} {\text {curl}}^{2D}{{\varvec{\psi }}^k_I} = \frac{ \alpha _I \beta _I }{{\text {det}}{{{\varvec{J}}}}} {\text {curl}}^{2D}_{{\varvec{\xi }}}{{\varvec{v}}}^k_I\,. \end{aligned}$$
(31)
The micro-distortion field \({{\varvec{P}}}\) is approximated by the vectorial dofs \({{\varvec{d}}}^P_I\) representing its tangential components at location \(I=1,\ldots ,n^P\). The micro-distortion field and its Curl are interpolated as
$$\begin{aligned} {{\varvec{P}}}_h = \sum _{I=1}^{n^P} {{\varvec{d}}}^P_I \otimes {\varvec{\psi }}^k_I \, , \quad {\text {Curl}}{{\varvec{P}}}_h = \sum _{I=1}^{n^P} {{\varvec{d}}}^P_I \otimes {\text {curl}}{{{\varvec{\psi }}}^k_I} \,. \end{aligned}$$
(32)
The non-vanishing components of the Curl operator of the micro-distortion field for the 2D case are obtained by
$$\begin{aligned} \left[ \begin{array}{c} {\text {curl}}^{2D}{{{\varvec{P}}}_h^{1}} \\ {\text {curl}}^{2D}{{{\varvec{P}}}_h^{2}} \\ \end{array}\right]= & {} \sum _{I=1}^{n^P} {{\varvec{d}}}^P_I {\text {curl}}^{2D}{{{\varvec{\psi }}}^k_I} \nonumber \\= & {} \left[ \begin{array}{c} \sum _{I=1}^{n^P} (d^P_I)_1 {\text {curl}}^{2D}{{{\varvec{\psi }}}^k_I} \\ \sum _{I=1}^{n^P} (d^P_I)_2 {\text {curl}}^{2D}{{{\varvec{\psi }}}^k_I} \\ \end{array}\right] .\nonumber \\ \end{aligned}$$
(33)

3.3 Implemented finite elements

In this work, we present four nodal-edge elements based on the formulation in Sect. 3.2 and two standard nodal elements based on Sect. 3.1. All implemented finite elements employ scalar quadratic shape functions of Lagrange-type for the displacement field approximation with the notation T2 for triangles and Q2 for quadrilaterals. The micro-distortion field is approximated using different formulations introduced in Sects. 3.1 and 3.2. For the standard nodal elements, Lagrange-type ansatz functions are used resulting in element types T2T1 (linear ansatz for \({{\varvec{P}}}\)) and T2T2 (quadratic ansatz for \({{\varvec{P}}}\)). Different nodal-edge elements are built utilizing first- and second-order Nédélec formulations with tangential-conforming shape functions denoted as NT1 and NT2 for triangular elements and QT1 and QT2 for quadrilateral elements. The micro-distortion dofs in the standard nodal elements T2T1 and T2T2 are tensorial with \(2 \times 2\) entries while the nodal-edge elements use vectorial dofs for the micro-distortion field which represent the tangential components. The full micro-distortion tensor is restored based on Equation (32). The used finite elements are depicted in the parameter space in Fig. 3.
The expected convergence rates of \(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) elements for the relaxed micromorphic model were discussed for anti-plane shear and 3D cases in [50, 51]. In a similar way for an element with a \(H^1({{\mathcal {B}}})\)-conforming formulation of order k for the displacement approximation and a first type Nédélec formulation of order k for the micro-distortion approximation, the solution converges with an optimal convergence rate k in \(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) norm which is defined in Appendix A. Therefore, we expect that the elements T2NT2 and Q2NQ2 achieve an optimal convergence rate of two, see [50, 51].

3.4 Patch-test

The patch-test, introduced in the 1960s, is used for the following purposes: i) to check the performance of finite element formulations violating continuity requirements ii) as a necessary condition which all finite elements have to satisfy iii) to find out simple programming errors iv) as an established tool to check the convergence order for any element type, see [57]. For elements which satisfy required continuity conditions, a patch-test checks the correct programming.
We design a patch-test for the relaxed micromorphic model. A domain \({{\mathcal {B}}}= [0,1] \times [0,1]\) is assumed. The isotropic case of the relaxed micromorphic model is considered with material parameters \( \lambda _{\text {micro}} = \mu _{\text {micro}} = \lambda _e = \mu _e = \mu = 1\), \(L_{{\mathrm {c}}}= 1\,\) and \(\mu _c=0\), c.f. [44, 47]. The solution is assumed to be known a priori (\( {{\overline{{{\varvec{u}}}}}} \) and \( {{\overline{{{\varvec{P}}}}}} = \nabla {{\overline{{{\varvec{u}}}}}}\)) and by solving the strong forms, the body moments and forces are defined. For the numerical setup, the boundary value problem is built by enforcing the derived body moments and forces with Dirichlet boundary conditions \({{\varvec{u}}}= {{\overline{{{\varvec{u}}}}}}\) and \({{\varvec{P}}}\cdot {\varvec{\tau }}= {{\overline{{{\varvec{P}}}}}} \cdot {\varvec{\tau }}\) on the whole boundary \(\partial {{\mathcal {B}}}\,\). Irregular meshes consisting of four quadrilaterals or triangles are used, see Fig. 4.
Simple patch-test: First, we analyze the case of a linear displacement and a constant micro-distortion field. The assumed solution with the related body-forces and body-moments reads
$$\begin{aligned} {{\overline{{{\varvec{u}}}}}} = \left( \begin{array}{c} x \\ y \end{array} \right) , \quad {{\overline{{{\varvec{P}}}}}} = \left( \begin{array}{c c } 1 &{} 0 \\ 0 &{} 1 \\ \end{array} \right) , \quad {{\overline{{{\varvec{f}}}}}} = {\mathbf{0}}, \quad {{\overline{{{\varvec{M}}}}}} = \left( \begin{array}{c c } 4 &{} 0 \\ 0 &{} 4 \\ \end{array} \right) \,.\nonumber \\ \end{aligned}$$
(34)
All elements must capture the solution exactly due to the polynomial order of the shape functions and which is what we obtain after testing.
Table 1
Material parameters of the second boundary problem, see Fig. 14
Material 1
Material 2
\( \lambda _{\text {micro}} = 555.55 , \quad \mu _{\text {micro}} = 833.33\)
\( \lambda _{\text {micro}} = 2777.78 , \quad \mu _{\text {micro}} = 4166.67\)
\(\lambda _e = 486.11, \quad \mu _e = 729.17\)
\(\lambda _e = 2430.555, \quad \mu _e = 3645.85\)
\(\mu _c = 0, \quad \mu =833.33\)
\(\mu _c = 0, \quad \mu =4166.67\)
\({\mathbb {L}}= {\mathbb {I}}, \quad L_{{\mathrm {c}}}\in \{0.001, 5, 1000\}\)
\({\mathbb {L}}= {\mathbb {I}}, \quad L_{{\mathrm {c}}}\in \{0.001, 5, 1000\}\)
Higher-order patch-test: Next, we analyze the case of a quadratic displacement and a linear micro-distortion field. The assumed solution with the related body-forces and body-moments reads
$$\begin{aligned}&{{\overline{{{\varvec{u}}}}}} = \left( \begin{array}{c} x^2 \\ y^2 \end{array} \right) , \quad {{\overline{{{\varvec{P}}}}}} = \left( \begin{array}{c c } 2 x &{} 0 \\ 0 &{} 2 y \\ \end{array} \right) , \nonumber \\&\quad {{\overline{{{\varvec{f}}}}}} = {\mathbf{0}}, \quad {{\overline{{{\varvec{M}}}}}} = \left( \begin{array}{c c } 6 x +2 y &{} 0 \\ 0 &{} 2 x+ 6 y \\ \end{array} \right) \,. \end{aligned}$$
(35)
Due to the polynomial order of the shape functions, the elements T2T1, T2T2, T2NT2 and Q2NQ2 are able to obtain the analytical solution (except for the typical computational inaccuracies). First-order Nédélec elements are unable to capture the analytical solution without numerical errors. Therefore, we use this test to study the convergence behavior of first-order Nédélec elements T2NT1 and Q2NQ1. The used structured and unstructured meshes as well as the convergence behavior are depicted in Figs. 5 and 6. The numerical solution converges to the analytical solution when refining the mesh.

4 Numerical examples

For our numerical examples, we assume isotropic material behavior which can be described by the set of material parameters \(\lambda _{\text {micro}},\mu _{\text {micro}},\lambda _e,\mu _e,\mu _c ,\mu \) and \(L_{{\mathrm {c}}}\), where \(\lambda _*\) and \(\mu _*\) denote the Lamé coefficients. Furthermore, we assume \({\mathbb {L}}\) as fourth-order identity tensor \({\mathbb {I}}\). Throughout all examples, we consider the Cosserat modulus \(\mu _c = 0\), cf. [35, 38], leading to symmetry of the force stress tensor. The simulations presented in this paper are performed within AceGen and AceFEM programs, which are developed and maintained by Jože Korelc (University of Ljubljana). The interested reader is referred to [18].

4.1 Discontinuous solution: convergence study

In this boundary value problem (bvp), we consider a homogeneous rectangular domain \({{\mathcal {B}}}\) with length \(l=2\) and height \(h=1\), see Fig. 7. We assume the following displacement and micro-distortion field
$$\begin{aligned} \begin{aligned} {{\overline{{{\varvec{u}}}}}} =&{\left\{ \begin{array}{ll} \left( \begin{array}{c} e^{y (1-x) } \\ e^{y^2 (1-x)} \end{array} \right) \quad {\text {for}} \quad x \le 1 \\ \\ \left( \begin{array}{c} e^{y (x-1) } \\ e^{y^2 (x-1)} \end{array} \right) \quad {\text {for}} \quad x> 1 \end{array}\right. } \,, \\ {{\overline{{{\varvec{P}}}}}} = \nabla {{\overline{{{\varvec{u}}}}}} =&{\left\{ \begin{array}{ll} \left( \begin{array}{c} - y e^{y (1-x) } \quad (1-x)e^{y (1-x)} \\ -y^2 e^{y^2 (1-x)} \quad 2y(1-x)e^{y^2 (1-x) } \end{array} \right) \quad &{}{\text {for}} \quad x \le 1 \\ \\ \left( \begin{array}{c} y e^{y (x-1) } \quad (x-1)e^{y (x-1)} \\ y^2 e^{y^2 (x-1)} \quad 2y(x-1)e^{y^2 (x-1) } \end{array} \right) \quad&{\text {for}} \quad x > 1 \end{array}\right. } \,, \\ \end{aligned} \end{aligned}$$
(36)
where the displacement field and tangential components of the micro-distortion \({{{\overline{{{\varvec{P}}}}}} \cdot {{\varvec{e}}}_2 = ({{\overline{P}}}_{12},{{\overline{P}}}_{22})^T}\) are continuous on an interface at \(x=1\) while the normal components \({{{\overline{{{\varvec{P}}}}}} \cdot {{\varvec{e}}}_1 = ({{\overline{P}}}_{11},{{\overline{P}}}_{21})^T}\) exhibit discontinuities. The isotropic case of the relaxed micromorphic model is considered with the material parameters \( \lambda _{\text {micro}} = \mu _{\text {micro}} = \lambda _e = \mu _e = \mu = 1\), \(L_{{\mathrm {c}}}= 1\,\) and \(\mu _c=0\). Solving the strong forms yields vanishing body forces while the body moments read
$$\begin{aligned} {{\overline{{{\varvec{M}}}}}} = \left( \begin{array}{c c} 3 {\overline{u}}_{1,1} + {\overline{u}}_{2,2} \quad {\overline{u}}_{1,2} + {\overline{u}}_{2,1} \\ {\overline{u}}_{1,2} + {\overline{u}}_{2,1} \quad {\overline{u}}_{1,1} + 3 {\overline{u}}_{2,2} \end{array} \right) \,. \end{aligned}$$
(37)
For the numerical setup, the boundary conditions are set as
$$\begin{aligned} {{\varvec{u}}}= {{\overline{{{\varvec{u}}}}}} \quad {\text {and}} \quad {{\varvec{P}}}\cdot {\varvec{\tau }}= {{\overline{{{\varvec{P}}}}}} \cdot {\varvec{\tau }}= \nabla {{\overline{{{\varvec{u}}}}}} \cdot {\varvec{\tau }}\quad {\text {on}} \quad \partial {{\mathcal {B}}}\,, \end{aligned}$$
(38)
where the calculated body moments and different meshes are used. We show in Fig. 8 the displacement and micro-distortion field obtained by 882 Q2NQ2 elements on a structured rectangular mesh (see Fig. 11a). The tangential components of the micro-distortion are continuous on the interface while the normal components exhibit a jump.
A convergence study of the component \(P_{11}\) along the inspection line \(y=0.5\) is plotted in Figs. 9 and 10. \(H^1({{\mathcal {B}}}) \times H^1({{\mathcal {B}}})\) elements, see Fig. 9, result in a continuous solution and cause a transition zone at the interface which needs to be resolved by increasing the mesh density tremendously in order to approximate the discontinuous solution at the interface. Meanwhile, the discontinuous solution of \(P_{11}\) can be captured by \(H^1({{\mathcal {B}}}) \times H({\text {curl}}, {{\mathcal {B}}})\) elements, see Fig. 10. The second-order Nédélec formulations in T2NT2 and Q2NQ2 achieve a numerical solution close to the analytical solution even with a coarse mesh while first-order Nédélec formulations in T2NT1 and Q2NQ1 require a denser mesh, because they only represent a constant micro-distortion field in each element.
We study the convergence rates of the implemented finite elements on three different meshes, see Figs. 11, 12 and 13. The mesh schemes are refined similarly to Figs. 5 and 6. Instead of plotting \(||\{{{\varvec{u}}},{{\varvec{P}}}\} - \{{{\overline{{{\varvec{u}}}}}},{{\overline{{{\varvec{P}}}}}}\} ||^2_{H^1({{\mathcal {B}}})\times H({\text {curl}},{{\mathcal {B}}})}\), see Appendix A, we show its individual components. These are the \(L^2\)-norms of the error of the displacement \({||{{\varvec{u}}}-{{\overline{{{\varvec{u}}}}}}||_{L^2}}\), its gradient \({||\nabla {{\varvec{u}}}- \nabla {{\overline{{{\varvec{u}}}}}}||_{L^2}}\), the micro-distortion \(||{{\varvec{P}}}-{{\overline{{{\varvec{P}}}}}}||_{L^2}\) and its Curl field \({||{\text {Curl}}{{\varvec{P}}}- {\text {Curl}}{{\overline{{{\varvec{P}}}}}}||_{L^2}}\). The analysis of the individual error norms is more targeted-oriented from an engineering point of view, since it treats the different energy terms separately. Second-order Nédélec elements T2NT2 and Q2NQ2 achieve convergence rates of three in the \(L^2\)-norm \(||{{\varvec{u}}}-{{\overline{{{\varvec{u}}}}}}||_{L^2}\), and a convergence rate of two in the remaining error norms. In the analyzed bvp the second-order Nédélec elements T2NT2 and Q2NQ2 lead to an optimal convergence rate of two in the space \(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) norm, which meets the theoretical expectations. First-order Nédélec elements T2NT1 and Q2NQ1 show one order of convergence less compared to second-order ones.

4.2 Characteristic length analysis: pure shear problem

We introduce a second boundary value problem, see Fig. 14, consisting of a circular domain \({{\mathcal {B}}}\) with radius \(r_{\text {o}}= 25\) and a circular hole at its center with radius \(r_{\text {i}} = 2\). No body forces or moments are considered. We fix the displacement field \( {\overline{{{\varvec{u}}}}} = {\mathbf{0}}\) on the inner boundary \(\partial {{\mathcal {B}}}_{\text {i}}\) and we rotate the outer boundary \(\partial {{\mathcal {B}}}_{\text {o}}\) counter clockwise with \( {\bar{{{\varvec{u}}}}} = (-\frac{\Delta }{r_{\text {o}}} y, \frac{\Delta }{r_{\text {o}}} x)^T\) where \(\Delta = 0.01\). For the micro-distortion field, we apply the consistent coupling boundary condition (\({{\varvec{P}}}\cdot {\varvec{\tau }}= \nabla {\overline{{{\varvec{u}}}}} \cdot {\varvec{\tau }}\)) on all boundaries \(\partial {{\mathcal {B}}}= \partial {{\mathcal {B}}}_{\text {i}} \cup \partial {{\mathcal {B}}}_{\text {o}} \). Two different cases are discussed in the following. For case A, a single material is assumed whereas for case B two materials are considered. The second material is located as a ring with an outer radius \(r_{\text {m}} = 10\) and an inner radius \(r_{\text {i}} = 2\). The material parameters are shown in Table 1. For the analysis of the influence of the characteristic length \(L_{{\mathrm {c}}}\), the characteristic length is varied.
The problem results in a rotationally-symmetric solution where only the shear components (\(u_{r,\theta }, u_{\theta ,r}, P_{r \theta }, P_{\theta r} \ne 0\)) are non-vanishing. The convergence behavior of the different elements is investigated for case B and \(L_{{\mathrm {c}}}= 5\) using three different mesh densities (410, 3044 and 30620 triangular elements and 448, 3040 and 30256 quadrilateral elements). Since the micro-distortion field is in \(H({\text {curl}},{{\mathcal {B}}})\), the tangential shear component \(P_{r \theta }\) has to be continuous while the radial shear component \(P_{\theta r}\) exhibits a jump, see Fig. 15, where Q2NQ2 elements are used. Similar to Sect. 4.1, \(H^1({{\mathcal {B}}}) \times H^1({{\mathcal {B}}})\) elements are unable to capture this discontinuity in \(P_{\theta r}\), which is shown in Fig. 16. The discontinuous solution of the micro-distortion field is demonstrated in Fig. 17 using \(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) elements. The higher-order Nédélec formulation in T2NT2 and Q2NQ2 elements exhibits very satisfactory results already with a coarse mesh.
In the following, we will analyze the influence of a variation of the length-scale parameter \(L_{{\mathrm {c}}}\) on the response of the relaxed micromorphic model for case A. The relation of the relaxed micromorphic model to the classical Cauchy theory has been discussed in detail in [3, 40] for the limiting case \(L_{{\mathrm {c}}}\rightarrow 0\) and \(L_{{\mathrm {c}}}\rightarrow \infty \). \(L_{{\mathrm {c}}}\rightarrow 0\) relates to a macroscopic view on the material with microstructure, with the relaxed micromorphic model being equivalent to a linear elasticity model with stiffness \({\mathbb {C}}_{{\mathrm {macro}}}\) defined as the Reuss lower-bound of \({\mathbb {C}}_{{\mathrm {e}}}\) and \({\mathbb {C}}_{{\mathrm {micro}}}\), i.e. \({\mathbb {C}}_{{\mathrm {macro}}}:= ({\mathbb {C}}_{{\mathrm {e}}}^{-1} + {\mathbb {C}}_{{\mathrm {micro}}}^{-1})^{-1}\). The case \(L_{{\mathrm {c}}}\rightarrow \infty \) resembles an infinite zoom into the material, where an equivalence to linear elasticity with \({\mathbb {C}}_{{\mathrm {micro}}}\) can be derived, cf. [40]. In the latter case, it can be shown that \({{\varvec{P}}}= \nabla {{\varvec{u}}}\) holds. In our study, we approximate the limiting cases by \(L_{{\mathrm {c}}}=10^{-3}\) and \(L_{{\mathrm {c}}}=10^{3}\), respectively. Figure 18 shows the elastic energy W along the radius and Fig. 19 illustrates the non-vanishing components of \({{\varvec{P}}}\) together with the respective displacement gradient components using 30624 Q2NQ2 elements. In Fig. 20, we plot the total potential of the relaxed micromorphic model while varying the characteristic length parameter \(L_{{\mathrm {c}}}\). The figures clearly show the above described behavior. Bounding behavior of the relaxed micromorphic model for small sizes is an important advantage which most other generalized continua miss. Nevertheless, the previous results do not hold for different boundary conditions of the micro-distortion field. Using a different setting of Dirichlet boundary conditions (e.g. homogeneous Dirichlet boundary condition) will keep the role of the characteristic length (increasing \(L_{{\mathrm {c}}}\) makes the material stiffer) but the upper limit, when \(L_{{\mathrm {c}}}\rightarrow \infty \), will be reliant on the boundary value problem and the applied boundary conditions. Using the consistent coupling boundary condition allows the model to reproduce Cauchy linear elasticity with \({\mathbb {C}}_{{\mathrm {micro}}}\) and \({{\varvec{P}}}= \nabla {{\varvec{u}}}\) for \(L_{{\mathrm {c}}}\rightarrow \infty \) regardless of the boundary value problem, see [4447]. For the consistent coupling boundary condition, \({\mathbb {C}}_{{\mathrm {micro}}}\) can be related to the stiffest response of the material on the smallest reasonable scale such as one unit cell of a metamaterial. For the linear elasticity model, a standard T2 nodal element is implemented.
Next, we investigate the behavior of the different stresses \({\varvec{\sigma }}\), \({\varvec{\sigma }}_{\text {micro}}\) and \({{\varvec{m}}}\) under a variation of \(L_{{\mathrm {c}}}\). The force stress tensor \({\varvec{\sigma }}\) shown in Fig. 21 vanishes for large value of the characteristic length, \(L_{{\mathrm {c}}}=1000\), while it is bounded from above by the classical linear elasticity stress with elasticity tensor \({\mathbb {C}}_{{\mathrm {macro}}}\) for \(L_{{\mathrm {c}}}=0.001\). The only non-vanishing component of the moment stress \(m_{r z}\) is shown in Fig. 22 (\(m_{\theta z} = 0\)), which behaves opposite to the force stress when varying \(L_{{\mathrm {c}}}\). It is nearly zero for \(L_{{\mathrm {c}}}=0.001\) and it increases for growing \(L_{{\mathrm {c}}}\). The micro-stress shown in Fig. 23 is confined between the linear elasticity stress with elasticity tensor \({\mathbb {C}}_{{\mathrm {micro}}}\) from above and \({\mathbb {C}}_{{\mathrm {macro}}}\) from below for large and small values of the characteristic length, respectively. Summarizing the previous findings shortly, increasing the characteristic length diminishes force stress and raises micro- and moment stresses, while both force and moment stresses vanish for large and small values of the characteristic length, respectively, micro-stress is always present.

5 Conclusions

The relaxed micromorphic model is a generalized continuum model which can suitably reproduce the macroscopic effective properties of mechanical metamaterials. First, we derived the variational problem with the relevant weak and strong forms and associated boundary conditions. We put together the main components of standard nodal and nodal-edge finite element formulations of the relaxed micromorphic model. The standard nodal elements \(H^1({{\mathcal {B}}}) \times H^1({{\mathcal {B}}})\) are incapable to achieve satisfactory results for a discontinuous solution. \(H^1({{\mathcal {B}}}) \times H({\text {curl}}, {{\mathcal {B}}})\) elements capture the jumps of the normal components of the micro-distortion field. In contrast to the standard nodal elements, we observe an efficient convergence in the sense of the error norm reduction with mesh refinement. We reveal the role of the characteristic length which governs the scale-dependency property of the relaxed micromorphic model. For \(L_{{\mathrm {c}}}\rightarrow 0\), the model is equivalent to the standard Cauchy linear elasticity model with \({\mathbb {C}}_{{\mathrm {macro}}}\) defined as the Reuss lower-limit of elasticity tensors \({\mathbb {C}}_{{\mathrm {e}}}\) and \({\mathbb {C}}_{{\mathrm {micro}}}\) while the model is corresponding to Cauchy linear elasticity model with \({\mathbb {C}}_{{\mathrm {micro}}}\) with \({{\varvec{P}}}= \nabla {{\varvec{u}}}\) for \(L_{{\mathrm {c}}}\rightarrow \infty \). Furthermore, we have shown the dependency of different stress measurements on the characteristic length. The force stress is at maximum for \(L_{{\mathrm {c}}}\rightarrow 0\) and it vanishes for \(L_{{\mathrm {c}}}\rightarrow \infty \) but the moment stress behaves in the opposite way. The micro-stress varies between Cauchy linear elasticity stresses with \({\mathbb {C}}_{{\mathrm {micro}}}\) and \({\mathbb {C}}_{{\mathrm {macro}}}\) for \(L_{{\mathrm {c}}}\rightarrow \infty \) and \(L_{{\mathrm {c}}}\rightarrow 0\), respectively.

Acknowledgements

Funded by the Deutsche Forschungsgemeinschaft (DFG, German research Foundation) - Project number 440935806 (SCHR 570/39-1, SCHE 2134/1-1, NE 902/10-1) within the DFG priority program 2256. The authors gratefully acknowledge Jože Korelc for the development and ongoing support when using AceGen and AceFEM.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Anhänge

\(H^1({{\mathcal {B}}}) \times H({\text {curl}},{{\mathcal {B}}})\) norm

In order to define the related norm of the space \(H^1({{\mathcal {B}}})\times H({\text {curl}},{{\mathcal {B}}})\), we introduce the Lebesgue space
$$\begin{aligned} L^2({{\mathcal {B}}}) = \{ a \,| \,\, ||a||_{L^2} < \infty \,, \quad ||a||^2_{L^2} = \mathop {\int }\limits _{{\mathcal {B}}}a^2 {\text {d}} V \} \,, \end{aligned}$$
(A.1)
and the Lebesgue norms of a vector-valued function \({{\varvec{a}}}\) and a second-order tensor \({{\varvec{A}}}\) read
$$\begin{aligned} ||{{\varvec{a}}}||^2_{L^2} = \sum _{i=1}^3 ||a_i||^2_{L^2} \,,\quad ||{{\varvec{A}}}||^2_{L^2} = \sum _{i=1}^3 \sum _{j=1}^3 ||A_{ij}||^2_{L^2} \,. \end{aligned}$$
(A.2)
The Hilbert space \(H^1({{\mathcal {B}}})\) is defined for \({{\varvec{u}}}\) as
$$\begin{aligned} H^1({{\mathcal {B}}}) = \{ {{\varvec{u}}}\in L^2({{\mathcal {B}}}) \,|\, \nabla {{\varvec{u}}}\in L^2({{\mathcal {B}}}) \} \, \end{aligned}$$
(A.3)
with the norm
$$\begin{aligned} ||{{\varvec{u}}}||^2_{H^1({{\mathcal {B}}})} = ||{{\varvec{u}}}||^2_{L^2} + ||\nabla {{\varvec{u}}}||^2_{L^2}\,. \end{aligned}$$
(A.4)
The space \(H({\text {curl}},{{\mathcal {B}}})\) is defined for \({{\varvec{P}}}\) as
$$\begin{aligned} H({\text {curl}},{{\mathcal {B}}})= & {} \{ {{\varvec{P}}}^i \in L^2({{\mathcal {B}}}) \,|\, {\text {curl}}{{\varvec{P}}}^i \in L^2({{\mathcal {B}}}) \,, \quad {\text {for}}\nonumber \\&\quad i=1,2,3 \} \,, \end{aligned}$$
(A.5)
where \({{\varvec{P}}}^i\) are the row-vectors of \({{\varvec{P}}}\), see Equations 13 and 14, and the associated norm reads
$$\begin{aligned} ||{{\varvec{P}}}||^2_{H({\text {curl}},{{\mathcal {B}}})} = ||{{\varvec{P}}}||^2_{L^2} + ||{\text {Curl}}{{\varvec{P}}}||^2_{L^2} \,. \end{aligned}$$
(A.6)
The norm of the space \(H^1({{\mathcal {B}}})\times H({\text {curl}},{{\mathcal {B}}})\) is defined as
$$\begin{aligned} ||\{{{\varvec{u}}},{{\varvec{P}}}\} ||^2_{H^1({{\mathcal {B}}})\times H({\text {curl}},{{\mathcal {B}}})} = ||{{\varvec{u}}}||^2_{H^1({{\mathcal {B}}})} + ||{{\varvec{P}}}||^2_{H({\text {curl}},{{\mathcal {B}}})} \,.\nonumber \\ \end{aligned}$$
(A.7)
For further details, we refer to [51].

Nédélec shape functions

Construction of triangular Nédélec shape functions

The implemented finite elements in the parameter space are defined with natural coordinates \({{\varvec{\xi }}= (\xi ,\eta )^T}\). Triangular elements are defined on a domain \({{\mathcal {B}}}_e^\triangle = \{0 \le \xi \le 1, 0 \le \eta \le 1 - \xi \}\). The finite elements with the respective edge numbering are shown in Fig. 1.

First-order triangular element NT1

The Nédélec space of a first-order triangular element (NT1) reads
$$\begin{aligned} \left[ \mathcal{ND}\mathcal{}^\triangle \right] ^{2}_1 = \bigg \{ \left[ \begin{array}{c} 1 \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ 1 \end{array} \right] \,, \left[ \begin{array}{c} - \eta \\ \xi \end{array} \right] \bigg \}\,, \end{aligned}$$
(B.1)
and the general form of vectorial shape function is
$$\begin{aligned} {{\varvec{v}}}^1 = \left[ \begin{array}{c} a_1 - a_3 \, \eta \\ a_2 + a_3 \, \xi \end{array}\right] , \end{aligned}$$
(B.2)
where \(a_i, i=1,2,3\) are coefficients yet to be defined based on the dofs. Starting from the definition in Equation (24), we set \(r_j =1\) for all edges. The tangential vectors of all edges, see Fig. 1 (right), are given by
$$\begin{aligned} {{\varvec{t}}}_1 = \frac{1}{\sqrt{2}}\left[ \begin{array}{c} 1 \\ -1 \\ \end{array} \right] , \quad {{\varvec{t}}}_2 = \left[ \begin{array}{c} 0 \\ 1 \\ \end{array} \right] , \quad {{\varvec{t}}}_3 = \left[ \begin{array}{c} 1 \\ 0 \\ \end{array} \right] . \end{aligned}$$
(B.3)
We calculate the dofs following Equation (24) using \(\xi + \eta = 1\) on the first edge, \(\xi =0\) on the second edge and \(\eta =0\) on the third edge and obtain
$$\begin{aligned} m^{e_1}_{1} = a_1- a_2 - a_3, \qquad m^{e_2}_{1} = a_2, \qquad m^{e_3}_{1} = a_1\,. \end{aligned}$$
(B.4)
In order to obtain the vectorial shape functions \( {{\varvec{v}}}^1_1, {{\varvec{v}}}^1_2\) and \( {{\varvec{v}}}^1_3\) from the general function in (B.2), we have to compute three associated combinations for \(a_1, a_2\) and \(a_3\). We derive the explicit expressions for the vectorial shape functions by enforcing
$$\begin{aligned} m^{e_i}_{1} ({{\varvec{v}}}_j^k) = \delta _{ij}\, \end{aligned}$$
(B.5)
for function \({{\varvec{v}}}_j^k\) associated with edge \(e_j\). The evaluation of dofs for each edge, i.e.,
$$\begin{aligned} \begin{aligned} {\text {edge 1:}} \quad&m^{e_1}_{1} = 1, \quad m^{e_2}_{1} = 0, m^{e_3}_{1} = 0 \\ {}&\quad \Rightarrow \quad a_1=0, \quad a_2=0, \quad a_3=-1 \\ {\text {edge 2:}} \quad&m^{e_1}_{1} = 0, \qquad m^{e_2}_{1} = 1, m^{e_3}_{1} = 0 \\&\quad \Rightarrow \quad a_1=0, \quad a_2=1, \quad a_3=-1 \\ {\text {edge 3:}} \quad&m^{e_1}_{1} = 0, \qquad m^{e_2}_{1} = 0, m^{e_3}_{1} = 1 \\&\quad \Rightarrow \quad a_1=1, \quad a_2=0, \quad a_3=\,\,\,\,1 \\ \end{aligned} \end{aligned}$$
(B.6)
leads to the shape vectors
$$\begin{aligned} {{\varvec{v}}}^1_1 = \left( \begin{array}{c} \eta \\ - \xi \end{array} \right) \,, \quad {{\varvec{v}}}^1_2 = \left( \begin{array}{c} \eta \\ 1 - \xi \end{array} \right) \,, \quad {{\varvec{v}}}^1_3 = \left( \begin{array}{c} 1 - \eta \\ \xi \end{array} \right) \,.\nonumber \\ \end{aligned}$$
(B.7)
A visualization is depicted in Fig. 24.

Second-order triangular element NT2

The Nédélec space of a second-order triangular element (NT2) reads
$$\begin{aligned} \left[ \mathcal{ND}\mathcal{}^\triangle \right] ^{2}_2&= \bigg \{ \left[ \begin{array}{c} 1 \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \xi \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \eta \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ 1 \end{array} \right] \,,\nonumber \\&\qquad \left[ \begin{array}{c} 0 \\ \xi \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \eta \end{array} \right] \,, \left[ \begin{array}{c} - \eta ^2 \\ \xi \eta \end{array} \right] \,, \left[ \begin{array}{c} - \xi \eta \\ \xi ^2 \end{array} \right] \bigg \}\,, \end{aligned}$$
(B.8)
and the general form of shape functions reads
$$\begin{aligned} {{\varvec{v}}}^2 = \left( \begin{array}{c} a_1 + a_2 \, \xi + a_3 \, \eta - a_7 \, \eta ^2 - a_8 \, \xi \eta \\ a_4 + a_5 \, \xi + a_6 \, \eta + a_7 \, \xi \eta + a_8 \, \xi ^2 \end{array} \right) , \end{aligned}$$
(B.9)
where \(a_i, i=1,\ldots ,8\) are coefficients yet to be defined based on the dofs. Explicit functions \({r_j, {j=1,2}}\) and \({{{\varvec{q}}}_i, { i=1,2}}\) in Equations (24) and (25) are assumed as
$$\begin{aligned} \begin{aligned}&{\text {edge 1 :}} \qquad&r_1 =\xi \,,&\qquad&r_2 = \eta \,, \\&{\text {edge 2 :}} \qquad&r_1 = \eta \,,&\qquad&r_2 = 1- \eta \,, \\&{\text {edge 3 :}} \qquad&r_1 = 1-\xi \,,&\qquad&r_2 = \xi \,, \\&{\text {inner :}} \qquad&{{\varvec{q}}}_1 = \left[ \begin{array}{c} 1 \\ 0 \end{array} \right] ,&\qquad&{{\varvec{q}}}_2 = \left[ \begin{array}{c} 0 \\ 1 \end{array} \right] , \end{aligned} \end{aligned}$$
(B.10)
and the tangential vectors of the element edges are the same as in Equation B.3 .
The inner and outer dofs are calculated according to Equations (24) and (25) using \(\xi + \eta = 1\) on the first edge, \(\xi =0\) on the second edge and \(\eta =0\) on the third edge, such as
$$\begin{aligned}&m^{e_1}_{1} = \frac{1}{6}(3 a_1 + 2 a_2 + a_3 - 3 a_4 - 2 a_5 - a_6 - a_7 - 2 a_8) \,,\nonumber \\&m^{e_1}_{2} =\frac{1}{6} (3 a_1 + a_2 + 2 a_3 - 3 a_4 - a_5 - 2 a_6 - 2 a_7 - a_8) \,,\nonumber \\&m^{e_2}_{1} = \frac{1}{6} (3a_4 + 2 a_6) \,, \qquad m^{e_2}_{2} = \frac{1}{6} ( 3a_4 + a_6) \,,\nonumber \\&m^{e_3}_{1} = \frac{1}{6} (3a_1 + a_2) \,, \qquad \,\,\, m^{e_3}_{2} = \frac{1}{6} (3a_1 + 2a_2) \,,\nonumber \\&m^{\text {inner}}_1 = \frac{1}{24} (12 a_1 + 4 a_2 + 4 a_3 - 2 a_7 - a_8) \,,\nonumber \\&m^{\text {inner}}_2 = \frac{1}{24} (12 a_4 + 4 a_5 + 4 a_6 + a_7 + 2 a_8) \,, \end{aligned}$$
(B.11)
and the resulting shape functions shown in Fig. 25 are obtained by an analogous procedure as described in B.1.1, leading to
$$\begin{aligned}&{\text {edge 1 :}} {{\varvec{v}}}_1^2= 2 \left( \begin{array}{c} -\eta + 4 \eta \xi \\ 2 \xi - 4 \xi ^2 \end{array} \right) ,\nonumber \\&\qquad \qquad {{\varvec{v}}}_2^2=2 \left( \begin{array}{c} -2 \eta + 4 \eta ^2 \\ \xi - 4 \eta \xi \end{array} \right) ,\nonumber \\&{\text {edge 2 :}} {{\varvec{v}}}_3^2 = 2\left( \begin{array}{c} -2 \eta + 4 \eta ^2 \\ -1 + 3 \eta + \xi - 4 \eta \xi \end{array} \right) ,\nonumber \\&\quad \qquad \quad {{\varvec{v}}}_4^2 = 2\left( \begin{array}{c} 3 \eta - 4 \eta ^2 - 4 \eta \xi \\ 2 - 3 \eta - 6 \xi + 4 \eta \xi + 4 \xi ^2 \end{array} \right) ,\nonumber \\&{\text {edge 3 :}} {{\varvec{v}}}_5^2 = 2\left( \begin{array}{c} 2 - 6 \eta + 4 \eta ^2 - 3 \xi + 4 \eta \xi \\ 3 \xi - 4 \eta \xi - 4 \xi ^2 \end{array} \right) ,\nonumber \\&\quad \qquad \quad {{\varvec{v}}}_6^2 = 2\left( \begin{array}{c} -1 + \eta + 3 \xi - 4 \eta \xi \\ -2 \xi + 4 \xi ^2 \end{array} \right) ,\nonumber \\&{\text {inner :}} {{\varvec{v}}}_7^2 = 2\left( \begin{array}{c} 8 \eta - 8 \eta ^2 - 4 \eta \xi \\ -4 \xi + 8 \eta \xi + 4 \xi ^2 \end{array} \right) ,\nonumber \\&\quad \qquad \quad {{\varvec{v}}}_8^2 = 2\left( \begin{array}{c} -4 \eta + 4 \eta ^2 + 8 \eta \xi \\ 8 \xi - 4 \eta \xi - 8 \xi ^2 \end{array} \right) . \end{aligned}$$
(B.12)

Construction of quadrilateral Nédélec shape functions

Quadrilateral elements have a domain \({{\mathcal {B}}}_e^\square = \{-1 \le \xi \le 1, -1 \le \eta \le 1\}\).

First-order quadrilateral element NQ1

The Nédélec space of a first-order quadrilateral element (NQ1) reads
$$\begin{aligned} \left[ \mathcal{ND}\mathcal{}^\square \right] ^{2}_1 = \bigg \{ \left[ \begin{array}{c} 1 \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \eta \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ 1 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \xi \end{array} \right] \bigg \} \,, \end{aligned}$$
(B.13)
and the general form of shape vectors reads
$$\begin{aligned} {{\varvec{v}}}^1 = \left( \begin{array}{c} a_1 + a_2 \, \eta \\ a_3 + a_4 \, \xi \end{array} \right) , \end{aligned}$$
(B.14)
where \(a_i, i=1,..,4\) are coefficients yet to be defined based on the dofs. Starting from Equation (24), we set \(r_j = 1\) for all edges. The tangential vectors for the first and third edge are \({{\varvec{t}}}_1 = {{\varvec{t}}}_3 = (1,0)^T\) and for the second and fourth edge \({{\varvec{t}}}_2 = {{\varvec{t}}}_4 = (0,1)^T\), see Fig. 1 (left). We calculate edge dofs taking into consideration that \(\eta = -1\) on the first edge, \(\xi = 1\) on the second edge, \(\eta = 1\) on the third edge and \(\xi = -1\) on the fourth edge, leading to
$$\begin{aligned}&m^{e_1}_{1} = 2(a_1 - a_2)\,, \quad m^{e_2}_{1} = 2(a_3 + a_4)\,, \quad m^{e_3}_{1} = 2(a_1 + a_2)\,, \nonumber \\&m^{e_4}_{1} = 2(a_3 - a_4) \,. \end{aligned}$$
(B.15)
We solve the system of equations obtained by an analogous procedure to Section B.1.1 leading to the shape functions demonstrated in Fig. 26 which read
$$\begin{aligned} \begin{aligned} {{\varvec{v}}}_{1}^1 =&\left( \begin{array}{c} (-\eta + 1)/4 \\ 0 \end{array} \right) , \quad {{\varvec{v}}}_{2}^1 = \left( \begin{array}{c} 0 \\ (\xi + 1)/4 \end{array} \right) , \\ {{\varvec{v}}}_{3}^1 =&\left( \begin{array}{c} (\eta + 1)/4 \\ 0 \end{array} \right) , \quad \quad {{\varvec{v}}}_{4}^1 = \left( \begin{array}{c} 0 \\ (-\xi + 1)/4 \end{array}\right) , \end{aligned} \end{aligned}$$
(B.16)
where \({{\varvec{v}}}^1_i\) is associated with edge \(e_i\) for \(i=1,..,4\) .

Second-order quadrilateral element NQ2

The Nédélec space of a second-order quadrilateral element (NQ2) reads
$$\begin{aligned} \left[ \mathcal{ND}\mathcal{}^\square \right] ^{2}_2 = \bigg \{&\left[ \begin{array}{c} 1 \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \xi \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \eta \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \xi \eta \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \eta ^2 \\ 0 \end{array} \right] \,, \left[ \begin{array}{c} \xi \eta ^2 \\ 0 \end{array} \right] \,, \nonumber \\&\left[ \begin{array}{c} 0 \\ 1 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \xi \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \eta \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \xi \eta \\ \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \xi ^2 \end{array} \right] \,, \left[ \begin{array}{c} 0 \\ \eta \xi ^2 \end{array} \right] \bigg \} \,,\nonumber \\ \end{aligned}$$
(B.17)
and vectorial shape functions have the following general form
$$\begin{aligned} {{\varvec{v}}}^2 = \left( \begin{array}{c} a_1 + a_2 \,\xi + a_3\, \eta + a_4\, \xi \eta + a_5\, \eta ^2 + a_6 \, \xi \eta ^2 \\ a_7 + a_8 \, \xi + a_9 \, \eta + a_{10} \, \xi \eta + a_{11} \, \xi ^2 + a_{12} \,\eta \xi ^2 \end{array} \right) \,,\nonumber \\ \end{aligned}$$
(B.18)
where \(a_i, i=1,..,12\) are coefficients yet to be defined based on the dofs. Starting from Equations (24) and (26), explicit functions \(r_j, {j=1,2}\) and \({{\varvec{q}}}_i, {i=1,2,3,4}\) are set as
$$\begin{aligned} {\text {edge 1:}} \quad r_1= & {} \frac{1}{2} (1- \xi )\,, \qquad r_2 =\frac{1}{2} (1+ \xi )\,, \nonumber \\ {\text {edge 2:}} \quad r_1= & {} \frac{1}{2} (1- \eta )\,, \qquad r_2 =\frac{1}{2} (1+ \eta )\,, \nonumber \\ {\text {edge 3:}} \quad r_1= & {} \frac{1}{2} (1+ \xi )\,, \qquad r_2 =\frac{1}{2} (1- \xi )\,, \nonumber \\ {\text {edge 4:}} \quad r_1= & {} \frac{1}{2} (1+ \eta )\,,\qquad r_2 =\frac{1}{2} (1- \eta )\,, \nonumber \\ {\text {inner :}} \quad {{\varvec{q}}}_1= & {} \left[ \begin{array}{c} \frac{1}{2} (1+ \xi ) \\ 0 \end{array} \right] , \quad {{\varvec{q}}}_2 =\left[ \begin{array}{c} \frac{1}{2} (1- \xi ) \\ 0 \end{array} \right] , \nonumber \\ {{\varvec{q}}}_3= & {} \left[ \begin{array}{c} 0 \\ \frac{1}{2} (1+ \eta ) \end{array} \right] , \qquad {{\varvec{q}}}_4 =\left[ \begin{array}{c} 0 \\ \frac{1}{2} (1- \eta ) \end{array} \right] \,.\nonumber \\ \end{aligned}$$
(B.19)
The edge and inner dofs are calculated according to Equations (24) and (26) considering that tangential vectors and coordinates coloration are identical to the NQ1 element, such that
$$\begin{aligned}&m^{e_1}_{1} = \frac{1}{3} \, (3a_1 - a_2 - 3a_3 + a_4 + 3a_5 - a_6) \,, \nonumber \\&m^{e_1}_{2} = \frac{1}{3} \, (3a_1 + a_2 - 3a_3 - a_4 + 3a_5 + a_6)\,,\nonumber \\&m^{e_2}_{1} = \frac{1}{3} \, (-a_{10} +3a_{11} - a_{12} + 3a_7 + 3a_8 - a_9) \,, \nonumber \\&m^{e_2}_{2} = \frac{1}{3} \, (a_{10} +3a_{11} + a_{12} + 3a_7 + 3a_8 + a_9) \,, \nonumber \\&m^{e_3}_{1} = \frac{1}{3} \,(3a_1 + a_2 + 3a_3 + a_4 + 3a_5 + a_6) \,, \nonumber \\&m^{e_3}_{2} = \frac{1}{3} \,(3a_1 - a_2 + 3a_3 - a_4 + 3a_5 - a_6) \,,\nonumber \\&m^{e_4}_{1} = \frac{1}{3} \,( -a_{10} + 3a_{11} + a_{12} + 3a_7 - 3a_8 + a_9) \,, \nonumber \\&m^{e_4}_{2} = \frac{1}{3} \,( a_{10} + 3a_{11} - a_{12} + 3a_7 - 3a_8 - a_9) \,, \nonumber \\&m^{\text {inner}}_1 = \frac{2}{9} (9 a_1 + 3 a_2 + 3 a_5 + a_6) \,,\nonumber \\&m^{\text {inner}}_2 = \frac{2}{9} (9 a_1 - 3 a_2 + 3 a_5 - a_6) \,, \nonumber \\&m^{\text {inner}}_3 = \frac{2}{9} (3 a_{11} + a_{12} + 9 a_7 + 3 a_9) \,, \nonumber \\&m^{\text {inner}}_4 = \frac{2}{9} (3 a_{11} - a_{12} + 9 a_7 - 3 a_9) \,. \end{aligned}$$
(B.20)
Tangential-conforming shape functions demonstrated in Fig. 27 are obtained by an analogous procedure as described in B.1.1, leading to
$$\begin{aligned} {\text {edge 1:}} \quad {{\varvec{v}}}^2_1&= \left( \begin{array}{c} -1/8 - \eta /4 + 3 \eta ^2 /8 + 3 \xi /8 + 3 \eta \xi /4 - 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) , \nonumber \\ {{\varvec{v}}}^2_2&= \left( \begin{array}{c} -1/8 - \eta /4 + 3 \eta ^2 /8 - 3 \xi /8 - 3 \eta \xi /4 + 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) , \nonumber \\ {\text {edge 2:}} \quad {{\varvec{v}}}^2_3&= \left( \begin{array}{c} 0 \\ - 1/8 + 3 \eta /8 + \xi /4 - 3 \eta \xi /4 + 3 \xi ^2/8 - 9 \eta \xi ^2/8 \end{array} \right) , \nonumber \\ {{\varvec{v}}}^2_4&= \left( \begin{array}{c} 0 \\ - 1/8 - 3 \eta /8 + \xi /4 + 3 \eta \xi /4 + 3 \xi ^2/8 + 9 \eta \xi ^2/8 \end{array} \right) , \nonumber \\ {\text {edge 3:}} \quad {{\varvec{v}}}^2_5&= \left( \begin{array}{c} -1/8 + \eta /4 + 3 \eta ^2/8 - 3 \xi /8 + 3 \eta \xi /4 + 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) , \nonumber \\ {{\varvec{v}}}^2_6&= \left( \begin{array}{c} -1/8+ \eta /4 + 3 \eta ^2/8 + 3 \xi /8 - 3 \eta \xi /4 - 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) , \nonumber \\ {\text {edge 4:}} \quad {{\varvec{v}}}^2_7&= \left( \begin{array}{c} 0 \\ -1/8 - 3 \eta /8 - \xi /4 - 3 \eta \xi /4 + 3 \xi ^2/8 + 9 \eta \xi ^2/8 \end{array} \right) , \nonumber \\ {{\varvec{v}}}^2_8&= \left( \begin{array}{c} 0 \\ -1/8 + 3 \eta /8 - \xi /4 + 3 \eta \xi /4 + 3 \xi ^2/8 - 9 \eta \xi ^2/8 \end{array} \right) \,, \nonumber \\ {\text {inner:}} \quad {{\varvec{v}}}^2_9&= \left( \begin{array}{c} 3/8 - 3 \eta ^2/8 + 9 \xi /8 - 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) \,, \nonumber \\ {{\varvec{v}}}^2_{10}&= \left( \begin{array}{c} 3/8 - 3 \eta ^2/8 - 9 \xi /8 + 9 \eta ^2 \xi /8 \\ 0 \end{array} \right) \,,\nonumber \\ {{\varvec{v}}}^2_{11}&= \left( \begin{array}{c} 0 \\ 3/8 +9 \eta /8 - 3 \xi ^2/8 - 9 \eta \xi ^2/8 \end{array} \right) \,,\nonumber \\ {{\varvec{v}}}^2_{12}&= \left( \begin{array}{c} 0 \\ 3/8 -9 \eta /8 - 3 \xi ^2/8 + 9 \eta \xi ^2/8 \end{array} \right) \,. \end{aligned}$$
(B.21)
Literatur
1.
Zurück zum Zitat Aivaliotis A, Tallarico D, d‘Agostino MV, Daouadji A, Neff P, Madeo A (2020) Frequency- and angle-dependent scattering of a finite-sized meta-structure via the relaxed micromorphic model. Arch Appl Mech 90:1073–1096CrossRef Aivaliotis A, Tallarico D, d‘Agostino MV, Daouadji A, Neff P, Madeo A (2020) Frequency- and angle-dependent scattering of a finite-sized meta-structure via the relaxed micromorphic model. Arch Appl Mech 90:1073–1096CrossRef
2.
Zurück zum Zitat Altan BS, Aifantis EC (1997) On some aspects in the special theory of gradient elasticity. J Mech Behav Mater 8(3):231–282CrossRef Altan BS, Aifantis EC (1997) On some aspects in the special theory of gradient elasticity. J Mech Behav Mater 8(3):231–282CrossRef
3.
Zurück zum Zitat Barbagallo G, Madeo A, d’Agostino MV, Abreu R, Ghiba I-D, Neff P (2017) Transparent anisotropy for the relaxed micromorphic model: Macroscopic consistency conditions and long wave length asymptotics. Int J Solids Struct 120:7–30CrossRef Barbagallo G, Madeo A, d’Agostino MV, Abreu R, Ghiba I-D, Neff P (2017) Transparent anisotropy for the relaxed micromorphic model: Macroscopic consistency conditions and long wave length asymptotics. Int J Solids Struct 120:7–30CrossRef
4.
Zurück zum Zitat Barbagallo G, Tallarico D, d’Agostino MV, Aivaliotis A, Neff P, Madeo A (2019) Relaxed micromorphic model of transient wave propagation in anisotropic band-gap metastructures. Int J Solids Struct 162:148–163CrossRef Barbagallo G, Tallarico D, d’Agostino MV, Aivaliotis A, Neff P, Madeo A (2019) Relaxed micromorphic model of transient wave propagation in anisotropic band-gap metastructures. Int J Solids Struct 162:148–163CrossRef
5.
Zurück zum Zitat Boffi D, Brezzi F, Fortin M (2014) Mixed Finite Element Methods and Applications, volume 44 of Springer Series in Computational Mathematics. Springer, Berlin, Heidelberg Boffi D, Brezzi F, Fortin M (2014) Mixed Finite Element Methods and Applications, volume 44 of Springer Series in Computational Mathematics. Springer, Berlin, Heidelberg
6.
Zurück zum Zitat Brezzi F, Douglas J, Marini LD (1985) Two families of mixed finite elements for second order elliptic problems. Numer Math 47:217–235MathSciNetMATHCrossRef Brezzi F, Douglas J, Marini LD (1985) Two families of mixed finite elements for second order elliptic problems. Numer Math 47:217–235MathSciNetMATHCrossRef
7.
Zurück zum Zitat Cosserat E, Cosserat F (1909) Theory of deformable bodies. Hermann and Sons Cosserat E, Cosserat F (1909) Theory of deformable bodies. Hermann and Sons
8.
Zurück zum Zitat Crowley CW, Silvester PP, Hurwitz H (1988) Covariant projection elements for 3D vector field problems. IEEE Trans Magn 24(1):397–400CrossRef Crowley CW, Silvester PP, Hurwitz H (1988) Covariant projection elements for 3D vector field problems. IEEE Trans Magn 24(1):397–400CrossRef
9.
Zurück zum Zitat d‘Agostino MV, Barbagallo G, Ghiba I-D, Eidel B, Neff P, Madeo A (2020) Effective description of anisotropic wave dispersion in mechanical band-gap metamaterials via the relaxed micromorphic model. J Elast 139:299–329MathSciNetMATHCrossRef d‘Agostino MV, Barbagallo G, Ghiba I-D, Eidel B, Neff P, Madeo A (2020) Effective description of anisotropic wave dispersion in mechanical band-gap metamaterials via the relaxed micromorphic model. J Elast 139:299–329MathSciNetMATHCrossRef
10.
Zurück zum Zitat d’Agostino MV, Rizzi G, Khan H, Lewintan P, Madeo A, Neff P (2021) The consistent coupling boundary condition for the classical micromorphic model: existence, uniqueness and interpretation of the parameters (submitted). arXiv:2112.12050 d’Agostino MV, Rizzi G, Khan H, Lewintan P, Madeo A, Neff P (2021) The consistent coupling boundary condition for the classical micromorphic model: existence, uniqueness and interpretation of the parameters (submitted). arXiv:​2112.​12050
11.
Zurück zum Zitat Eringen AC (1968) Mechanics of micromorphic continua. In: Mechanics of Generalized Continua, pages 18–35. Springer, Berlin, Heidelberg Eringen AC (1968) Mechanics of micromorphic continua. In: Mechanics of Generalized Continua, pages 18–35. Springer, Berlin, Heidelberg
13.
Zurück zum Zitat Fischer P, Klassen M, Mergheim J, Steinmann P, Müller R (2011) Isogeometric analysis of 2D gradient elasticity. Comput Mech 47:1432–0924 MathSciNetMATHCrossRef Fischer P, Klassen M, Mergheim J, Steinmann P, Müller R (2011) Isogeometric analysis of 2D gradient elasticity. Comput Mech 47:1432–0924 MathSciNetMATHCrossRef
14.
Zurück zum Zitat Fischer SCL, Hillen L, Eberl C (2020) Mechanical metamaterials on the way from laboratory scale to industrial applications: Challenges for characterization and scalability. Materials 13(16):3605CrossRef Fischer SCL, Hillen L, Eberl C (2020) Mechanical metamaterials on the way from laboratory scale to industrial applications: Challenges for characterization and scalability. Materials 13(16):3605CrossRef
15.
Zurück zum Zitat Ghiba I-D, Neff P, Madeo A, Placidi L, Rosi G (2015) The relaxed linear micromorphic continuum: existence, uniqueness and continuous dependence in dynamics. Math Mech Solids 20(10):1171–1197MathSciNetMATHCrossRef Ghiba I-D, Neff P, Madeo A, Placidi L, Rosi G (2015) The relaxed linear micromorphic continuum: existence, uniqueness and continuous dependence in dynamics. Math Mech Solids 20(10):1171–1197MathSciNetMATHCrossRef
16.
Zurück zum Zitat Jiang Y, Li Y (2018) 3D printed auxetic mechanical metamaterial with chiral cells and re-entrant cores. Sci Rep 8(1):2397CrossRef Jiang Y, Li Y (2018) 3D printed auxetic mechanical metamaterial with chiral cells and re-entrant cores. Sci Rep 8(1):2397CrossRef
17.
Zurück zum Zitat Kirby RC, Logg A, Rognes ME, Terrel AR (2012) Common and unusual finite elements. In: Automated Solution of Differential Equations by the Finite Element Method: The FEniCS Book, pages 95–119. Springer Berlin Heidelberg, Berlin, Heidelberg Kirby RC, Logg A, Rognes ME, Terrel AR (2012) Common and unusual finite elements. In: Automated Solution of Differential Equations by the Finite Element Method: The FEniCS Book, pages 95–119. Springer Berlin Heidelberg, Berlin, Heidelberg
18.
Zurück zum Zitat Korelc J, Wriggers P (2016) Automation of Finite Element Methods. Springer International Publishing Korelc J, Wriggers P (2016) Automation of Finite Element Methods. Springer International Publishing
19.
Zurück zum Zitat Lee J-H, Singer JP, Thomas EL (2012) Micro-/nanostructured mechanical metamaterials. Adv Mater 24(36):4782–4810CrossRef Lee J-H, Singer JP, Thomas EL (2012) Micro-/nanostructured mechanical metamaterials. Adv Mater 24(36):4782–4810CrossRef
20.
Zurück zum Zitat Lei M, Hong W, Zhao Z, Hamel C, Chen M, Lu H, Qi HJ (2019) 3D printing of auxetic metamaterials with digitally reprogrammable shape. ACS Applied Materials & Interfaces 11(25):22768–22776CrossRef Lei M, Hong W, Zhao Z, Hamel C, Chen M, Lu H, Qi HJ (2019) 3D printing of auxetic metamaterials with digitally reprogrammable shape. ACS Applied Materials & Interfaces 11(25):22768–22776CrossRef
21.
Zurück zum Zitat Leismann T, Mahnken R (2015) Comparison of hyperelastic micromorphic, micropolar and microstrain continua. Int J Non-Linear Mech 77:115–127CrossRef Leismann T, Mahnken R (2015) Comparison of hyperelastic micromorphic, micropolar and microstrain continua. Int J Non-Linear Mech 77:115–127CrossRef
22.
Zurück zum Zitat Leismann T, Mahnken R (2015) Transition from hyperelastic micromorphic to micropolar and microstrain continua. PAMM 15(1):329–330CrossRef Leismann T, Mahnken R (2015) Transition from hyperelastic micromorphic to micropolar and microstrain continua. PAMM 15(1):329–330CrossRef
23.
Zurück zum Zitat Madeo A, Neff P, Ghiba I-D, Placidi L, Rosi G (2015) Band gaps in the relaxed linear micromorphic continuum. Z Angew Math Mech 95(9):880–887MathSciNetMATHCrossRef Madeo A, Neff P, Ghiba I-D, Placidi L, Rosi G (2015) Band gaps in the relaxed linear micromorphic continuum. Z Angew Math Mech 95(9):880–887MathSciNetMATHCrossRef
24.
Zurück zum Zitat Madeo A, Neff P, Ghiba I-D, Placidi L, Rosi G (2015) Wave propagation in relaxed micromorphic continua: modeling metamaterials with frequency band-gaps. Continuum Mech Thermodyn 75:551–571MathSciNetMATHCrossRef Madeo A, Neff P, Ghiba I-D, Placidi L, Rosi G (2015) Wave propagation in relaxed micromorphic continua: modeling metamaterials with frequency band-gaps. Continuum Mech Thermodyn 75:551–571MathSciNetMATHCrossRef
25.
Zurück zum Zitat Madeo A, Neff P, d’Agostino MV, Barbagallo G (2016) Complete band gaps including non-local effects occur only in the relaxed micromorphic model. Comptes Rendus Mécanique 344(11–12):784–796CrossRef Madeo A, Neff P, d’Agostino MV, Barbagallo G (2016) Complete band gaps including non-local effects occur only in the relaxed micromorphic model. Comptes Rendus Mécanique 344(11–12):784–796CrossRef
26.
Zurück zum Zitat Madeo A, Neff P, Ghiba I-D, Rosi G (2016) Reflection and transmission of elastic waves in non-local band-gap metamaterials: a comprehensive study via the relaxed micromorphic model. J Mech Phys Solids 95:441–479MathSciNetMATHCrossRef Madeo A, Neff P, Ghiba I-D, Rosi G (2016) Reflection and transmission of elastic waves in non-local band-gap metamaterials: a comprehensive study via the relaxed micromorphic model. J Mech Phys Solids 95:441–479MathSciNetMATHCrossRef
27.
Zurück zum Zitat Madeo A, Neff P, Barbagallo G, d’Agostino MV, Ghiba I-D (2017) A review on wave propagation modeling in band-gap metamaterials via enriched continuum models. In: Mathematical Modelling in Solid Mechanics, volume 69 of Adv. Struct. Mater., pages 89–105. Springer, Singapore Madeo A, Neff P, Barbagallo G, d’Agostino MV, Ghiba I-D (2017) A review on wave propagation modeling in band-gap metamaterials via enriched continuum models. In: Mathematical Modelling in Solid Mechanics, volume 69 of Adv. Struct. Mater., pages 89–105. Springer, Singapore
28.
Zurück zum Zitat Madeo A, Collet M, Miniaci M, Billon K, Ouisse M, Neff P (2018) Modeling phononic crystals via the weighted relaxed micromorphic model with free and gradient micro-inertia. J Elast 130(1):59–83MathSciNetMATHCrossRef Madeo A, Collet M, Miniaci M, Billon K, Ouisse M, Neff P (2018) Modeling phononic crystals via the weighted relaxed micromorphic model with free and gradient micro-inertia. J Elast 130(1):59–83MathSciNetMATHCrossRef
30.
Zurück zum Zitat Mindlin RD, Eshel NN (1968) On first strain-gradient theories in linear elasticity. Int J Solids Struct 4(1):109–124 (ISSN 0020-7683)MATHCrossRef Mindlin RD, Eshel NN (1968) On first strain-gradient theories in linear elasticity. Int J Solids Struct 4(1):109–124 (ISSN 0020-7683)MATHCrossRef
31.
Zurück zum Zitat Monk P (1993) An analysis of Nédélec’s method for the spatial discretization of maxwell’s equations. J Comput Appl Math 47(1):101–121MathSciNetMATHCrossRef Monk P (1993) An analysis of Nédélec’s method for the spatial discretization of maxwell’s equations. J Comput Appl Math 47(1):101–121MathSciNetMATHCrossRef
32.
Zurück zum Zitat Montgomery SM, Kuang X, Armstrong CD, Qi HJ (2020) Recent advances in additive manufacturing of active mechanical metamaterials. Curr Opin Solid State Mater Sci 24(5):100869CrossRef Montgomery SM, Kuang X, Armstrong CD, Qi HJ (2020) Recent advances in additive manufacturing of active mechanical metamaterials. Curr Opin Solid State Mater Sci 24(5):100869CrossRef
35.
Zurück zum Zitat Neff P (2006) The Cosserat couple modulus for continuous solids is zero viz the linearized cauchy-stress tensor is symmetric. ZAMM - Journal of Applied Mathematics and Mechanics / Zeitschrift für Angewandte Mathematik und Mechanik 86(11):892–912MathSciNetMATHCrossRef Neff P (2006) The Cosserat couple modulus for continuous solids is zero viz the linearized cauchy-stress tensor is symmetric. ZAMM - Journal of Applied Mathematics and Mechanics / Zeitschrift für Angewandte Mathematik und Mechanik 86(11):892–912MathSciNetMATHCrossRef
36.
Zurück zum Zitat Neff P, Jeong J, Münch I, Ramézani H (2010) Linear Cosserat elasticity, conformal curvature and bounded stiffness. In: Mechanics of Generalized Continua: One Hundred Years After the Cosserats, pages 55–63. Springer New York Neff P, Jeong J, Münch I, Ramézani H (2010) Linear Cosserat elasticity, conformal curvature and bounded stiffness. In: Mechanics of Generalized Continua: One Hundred Years After the Cosserats, pages 55–63. Springer New York
37.
Zurück zum Zitat Neff P, Ghiba I-D, Madeo A, Placidi L, Rosi G (2014) A unifying perspective: the relaxed linear micromorphic continuum. Continuum Mech Thermodyn 26(5):639–681MathSciNetMATHCrossRef Neff P, Ghiba I-D, Madeo A, Placidi L, Rosi G (2014) A unifying perspective: the relaxed linear micromorphic continuum. Continuum Mech Thermodyn 26(5):639–681MathSciNetMATHCrossRef
38.
Zurück zum Zitat Neff P, Ghiba I-D, Lazar M, Madeo A (2015) The relaxed linear micromorphic continuum: well-posedness of the static problem and relations to the gauge theory of dislocations. The Quarterly Journal of Mechanics and Applied Mathematics 68(1):53–84MathSciNetMATHCrossRef Neff P, Ghiba I-D, Lazar M, Madeo A (2015) The relaxed linear micromorphic continuum: well-posedness of the static problem and relations to the gauge theory of dislocations. The Quarterly Journal of Mechanics and Applied Mathematics 68(1):53–84MathSciNetMATHCrossRef
39.
Zurück zum Zitat Neff P, Madeo A, Barbagallo G, d’Agostino MV, Abreu R, Ghiba I-D (2017) Real wave propagation in the isotropic-relaxed micromorphic model. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 473(2197):20160790MathSciNetMATHCrossRef Neff P, Madeo A, Barbagallo G, d’Agostino MV, Abreu R, Ghiba I-D (2017) Real wave propagation in the isotropic-relaxed micromorphic model. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 473(2197):20160790MathSciNetMATHCrossRef
40.
Zurück zum Zitat Neff P, Eidel B, d‘Agostino MV, Madeo A (2020) Identification of scale-independent material parameters in the relaxed micromorphic model through model-adapted first order homogenization. J Elast 139:269–298MathSciNetMATHCrossRef Neff P, Eidel B, d‘Agostino MV, Madeo A (2020) Identification of scale-independent material parameters in the relaxed micromorphic model through model-adapted first order homogenization. J Elast 139:269–298MathSciNetMATHCrossRef
41.
Zurück zum Zitat Olm M, Badia S, Martín AF (2019) On a general implementation of h- and p-adaptive curl-conforming finite elements. Adv Eng Softw 132:74–91CrossRef Olm M, Badia S, Martín AF (2019) On a general implementation of h- and p-adaptive curl-conforming finite elements. Adv Eng Softw 132:74–91CrossRef
42.
Zurück zum Zitat Plocher J, Panesar A (2019) Review on design and structural optimisation in additive manufacturing: Towards next-generation lightweight structures. Mater Des 183:108164CrossRef Plocher J, Panesar A (2019) Review on design and structural optimisation in additive manufacturing: Towards next-generation lightweight structures. Mater Des 183:108164CrossRef
43.
Zurück zum Zitat Raviart PA, Thomas JM (1977) A mixed finite element method for 2-nd order elliptic problems. In: Galligani I, Magenes E (eds) Mathematical Aspects of Finite Element Methods, pages 292–315, Berlin, Heidelberg. Springer Berlin Heidelberg Raviart PA, Thomas JM (1977) A mixed finite element method for 2-nd order elliptic problems. In: Galligani I, Magenes E (eds) Mathematical Aspects of Finite Element Methods, pages 292–315, Berlin, Heidelberg. Springer Berlin Heidelberg
44.
Zurück zum Zitat Rizzi G, Hütter G, Khan H, Ghiba I-D, Madeo A, Neff P (2021a) Analytical solution of the cylindrical torsion problem for the relaxed micromorphic continuum and other generalized continua (including full derivations). to appear in Mathematics and Mechanics of Solids Rizzi G, Hütter G, Khan H, Ghiba I-D, Madeo A, Neff P (2021a) Analytical solution of the cylindrical torsion problem for the relaxed micromorphic continuum and other generalized continua (including full derivations). to appear in Mathematics and Mechanics of Solids
45.
Zurück zum Zitat Rizzi G, Hütter G, Madeo A, Neff P (2021) Analytical solutions of the simple shear problem for micromorphic models and other generalized continua. Arch Appl Mech 91:2237–2254CrossRef Rizzi G, Hütter G, Madeo A, Neff P (2021) Analytical solutions of the simple shear problem for micromorphic models and other generalized continua. Arch Appl Mech 91:2237–2254CrossRef
46.
Zurück zum Zitat Rizzi G, Hütter G, Madeo A, Neff P (2021) Analytical solutions of the cylindrical bending problem for the relaxed micromorphic continuum and other generalized continua. Continuum Mech Thermodyn 33:1505–1539MathSciNetCrossRef Rizzi G, Hütter G, Madeo A, Neff P (2021) Analytical solutions of the cylindrical bending problem for the relaxed micromorphic continuum and other generalized continua. Continuum Mech Thermodyn 33:1505–1539MathSciNetCrossRef
47.
Zurück zum Zitat Rizzi G, Khan H, Ghiba I-D, Madeo A, Neff P (2021d) Analytical solution of the uniaxial extension problem for the relaxed micromorphic continuum and other generalized continua (including full derivations). to appear in Archive of Applied Mechanics Rizzi G, Khan H, Ghiba I-D, Madeo A, Neff P (2021d) Analytical solution of the uniaxial extension problem for the relaxed micromorphic continuum and other generalized continua (including full derivations). to appear in Archive of Applied Mechanics
48.
Zurück zum Zitat Rognes ME, Kirby RC, Logg A (2009) Efficient assembly of \(H({{\rm div}})\) and \(H({{\rm curl}})\) conforming finite elements. SIAM J Sci Comput 31(9):4130–4151MathSciNetMATH Rognes ME, Kirby RC, Logg A (2009) Efficient assembly of \(H({{\rm div}})\) and \(H({{\rm curl}})\) conforming finite elements. SIAM J Sci Comput 31(9):4130–4151MathSciNetMATH
49.
Zurück zum Zitat Schöberl J, Zaglmayr S (2005) High order Nédélec elements with local complete sequence properties. The International Journal for Computation and Mathematics in Electrical and Electronic Engineering 24(2):374–384MathSciNetMATHCrossRef Schöberl J, Zaglmayr S (2005) High order Nédélec elements with local complete sequence properties. The International Journal for Computation and Mathematics in Electrical and Electronic Engineering 24(2):374–384MathSciNetMATHCrossRef
50.
Zurück zum Zitat Sky A, Neunteufel M, Münch I, Schöberl J, Neff P (2021) A hybrid \(H^1\times H({\rm curl})\) finite element formulation for a relaxed micromorphic continuum model of antiplane shear. Comput Mech 68:1–24MathSciNetMATHCrossRef Sky A, Neunteufel M, Münch I, Schöberl J, Neff P (2021) A hybrid \(H^1\times H({\rm curl})\) finite element formulation for a relaxed micromorphic continuum model of antiplane shear. Comput Mech 68:1–24MathSciNetMATHCrossRef
51.
Zurück zum Zitat Sky A, Neunteufel M, Muench I, Schöberl J, Neff P (2022) Primal and mixed finite element formulations for the relaxed micromorphic model (submitted). arXiv:2202.08715 Sky A, Neunteufel M, Muench I, Schöberl J, Neff P (2022) Primal and mixed finite element formulations for the relaxed micromorphic model (submitted). arXiv:​2202.​08715
53.
Zurück zum Zitat Surjadi JU, Gao L, Du H, Li X, Xiong X, Fang NX, Lu Y (2019) Mechanical metamaterials and their engineering applications. Adv Eng Mater 21(3):1800864CrossRef Surjadi JU, Gao L, Du H, Li X, Xiong X, Fang NX, Lu Y (2019) Mechanical metamaterials and their engineering applications. Adv Eng Mater 21(3):1800864CrossRef
54.
Zurück zum Zitat Wriggers P (2008) Nonlinear Finite Element Methods. Springer, Berlin, HeidelbergMATH Wriggers P (2008) Nonlinear Finite Element Methods. Springer, Berlin, HeidelbergMATH
55.
Zurück zum Zitat Yu X, Zhou J, Liang H, Jiang Z, Wu L (2018) Mechanical metamaterials associated with stiffness, rigidity and compressibility: A brief review. Prog Mater Sci 94:114–173CrossRef Yu X, Zhou J, Liang H, Jiang Z, Wu L (2018) Mechanical metamaterials associated with stiffness, rigidity and compressibility: A brief review. Prog Mater Sci 94:114–173CrossRef
56.
57.
Zurück zum Zitat Zienkiewicz OC, Taylor RL (1997) The finite element patch test revisited a computer test for convergence, validation and error estimates. Comput Methods Appl Mech Eng 149(1):223–254MathSciNetMATHCrossRef Zienkiewicz OC, Taylor RL (1997) The finite element patch test revisited a computer test for convergence, validation and error estimates. Comput Methods Appl Mech Eng 149(1):223–254MathSciNetMATHCrossRef
Metadaten
Titel
Lagrange and based finite element formulations for the relaxed micromorphic model
verfasst von
Jörg Schröder
Mohammad Sarhil
Lisa Scheunemann
Patrizio Neff
Publikationsdatum
11.08.2022
Verlag
Springer Berlin Heidelberg
Erschienen in
Computational Mechanics / Ausgabe 6/2022
Print ISSN: 0178-7675
Elektronische ISSN: 1432-0924
DOI
https://doi.org/10.1007/s00466-022-02198-3

Weitere Artikel der Ausgabe 6/2022

Computational Mechanics 6/2022 Zur Ausgabe

Neuer Inhalt