In this section, the Kirchhoff–Love and the Reissner–Mindlin shell theory is recalled and the finite element formulation is given briefly.
The 3D shell problem
We start with the statement of the 3D shell problem, which is a 3D problem of linearized elasticity on special domains \(\varOmega\).
Find \({\tilde{{\mathbf u}}}: \varOmega \subset {\mathbb{R}}^3 \rightarrow {\mathbb{R}}^3\):
$$\begin{aligned} \begin{aligned} \nabla \cdot ({\tilde{\pmb {\sigma }}})+{\tilde{{\mathbf b}}}&=0 \quad&\text {in}\quad \varOmega ,\\ {\tilde{\pmb {\sigma }}}&= {\tilde{\mathbb {C}}} : {\tilde{\pmb {\epsilon }}}\quad&\text {in}\quad \varOmega , \\ {\tilde{\pmb {\epsilon }}}&= \frac{1}{2}\left( {\mathrm {grad}}( {\tilde{{\mathbf u}}})+{\mathrm {grad}}( {\tilde{{\mathbf u}}})^\top \right) \quad&\text {in}\quad \varOmega , \\ {\tilde{{\mathbf u}}}&= {\tilde{{\mathbf u}}}_D \quad&\text {on}\quad \varGamma _D, \\ {\tilde{{\mathbf t}}}&= {\tilde{{\mathbf t}}}_N \quad&\text {on}\quad \varGamma _N. \end{aligned} \end{aligned}$$
(1)
We assume that the domain \(\varOmega\) is defined through a parametrization \(\overline{\mathfrak {g}}\) of a reference surface \(\bar{\varOmega }\) (see Fig. 1).
$$\begin{aligned} \begin{aligned} \overline{\mathfrak {g}}:\;{\bar{U}}\subset {\mathbb{R}}^2&\rightarrow {\bar{\varOmega }}\subset {\mathbb{R}}^3 \\ (\theta ^1,\theta ^2)&\mapsto \overline{\mathfrak {g}}(\theta ^1,\theta ^2) .\end{aligned}\end{aligned}$$
(2)
In the present paper, we restrict us to the case of rectangular parameter domains \(\bar{\varOmega }\). Then the parametrization of \(\varOmega\) is given by
$$\begin{aligned} \begin{aligned} \mathfrak {g}:\; ( {{\bar{U}}\times T} ) \subset {\mathbb{R}}^3&\rightarrow \varOmega \subset {\mathbb{R}}^3 \\ (\theta ^1,\theta ^2)\times \theta ^3&\mapsto \mathfrak {g}(\theta ^1,\theta ^2,\theta ^3) = \overline{\mathfrak {g}}(\theta ^1,\theta ^2) + \theta ^3 \; \mathbf {n},\end{aligned}\end{aligned}$$
(3)
with the unit normal vector \(\mathbf {n}\) and the thickness interval \(T=\left[-\frac{t}{2},\frac{t}{2}\right]\), where t is the thickness of the shell. In the rest of the paper, Latin indices \(i,j,\ldots\) take the values 1, 2, 3 whereas Greek indices \(\alpha , \beta\) take the values 1, 2. Furthermore, Einstein summation convention applies. In the following, we need the covariant base vectors \(\mathbf {\overline{G}}_{\alpha }= \frac{\partial \bar{g} }{\partial \theta ^\alpha }\), \(\mathbf {G}_{\alpha }= \frac{\partial g }{\partial \theta ^\alpha }\), the covariant coefficients of the metric \(\overline{G}_{\alpha \beta } = \mathbf {\overline{G}}_{\alpha }\cdot \mathbf {\overline{G}}_{\beta }\), \(\mathrm G_{ij} = \mathbf {G}_{i} \cdot \mathbf {G}_{j}\), the contravariant coefficients of the metric \(\overline{G}^{\alpha \beta }=[\overline{G}_{\alpha \beta }]^{-1}\), \(\mathrm G^{\alpha \beta }=[\mathrm G_{\alpha \beta }]^{-1}\), and the contravariant base vectors \(\mathbf {\overline{G}}^{\alpha }= \overline{G}^{\alpha \beta } \mathbf {\overline{G}}_{\beta }\), \(\mathbf {G}^{i}= \mathrm G^{ij} \mathbf {G}_{j}\). Note that all quantities with a bar are defined on the reference surface and are, therefore, independent of \(\theta ^3\). Whereas quantities with no bar are defined on the three-dimensional shell volume. Instead of solving for \({\tilde{{\mathbf u}}}(\mathbf x),\mathbf x \in \varOmega\), we seek the displacement field \({\mathbf u}(\theta ^1,\theta ^2,\theta ^3) = {\tilde{{\mathbf u}}}(\mathfrak {g}(\theta ^1,\theta ^2,\theta ^3))\) defined on the parametric space. We assume that the bodyforce \({\mathbf b}\) is given with respect to a fixed Cartesian frame \({\mathbf {e}_{i}}\). Thus, \({\mathbf b}=b^i\mathbf {e}_{i}\) holds. Then, the balance of momentum in (1) in parametric coordinates reads
$$\begin{aligned} \left( \sigma ^{jl}_{,j} + \sigma ^{kl} \varGamma _{kj}^{\;\;\;\; j} + \sigma ^{jk} \varGamma _{kj}^{\;\;\;\; l} \right) J_l^i + b^i = 0, \end{aligned}$$
(4)
with the contravariant components of the stress tensor \(\pmb {\sigma }= \sigma ^{ij}\; \mathbf {G}_{i} \otimes \mathbf {G}_{j}\), the Christoffel symbols of second kind \(\varGamma _{ij}^{\;\;\;\; k} = \mathbf {G}^{k} _{\;,i} \cdot \mathbf {G}_{j}\), and \(J_l^i = \mathbf {G}_{l} \cdot \mathbf {e}^{i}\). Here and in the following the notation \(()_{,j}=\frac{\partial () }{\partial \theta ^i}\) applies. In the present paper, we employ a linear isotropic material law, where the contravariant components of the elasticity tensor \(\mathbb {C}=\mathbb {C}^{ijkl}\;\mathbf {G}_{i} \otimes \mathbf {G}_{j} \otimes \mathbf {G}_{k} \otimes \mathbf {G}_{l}\) are given by
$$\begin{aligned} \begin{aligned} \mathbb {C}^{\alpha \beta \gamma \varphi }&=\lambda \mathrm G^{\alpha \beta }\mathrm G^{\gamma \varphi }+\mu \left( \mathrm G^{\alpha \gamma }\mathrm G^{\beta \varphi }+\mathrm G^{\alpha \varphi }\mathrm G^{\beta \gamma }\right) ,\\ \mathbb {C}^{\alpha \beta 33}&= \mathbb {C}^{33\alpha \beta }= \lambda \mathrm G^{\alpha \beta },\\ \mathbb {C}^{3\alpha 3\beta }&=\mathbb {C}^{3\alpha \beta 3} =\mathbb {C}^{\alpha 33\beta }=\mathbb {C}^{\alpha 3\beta 3} = \mu \mathrm G^{\alpha \beta },\\ \mathbb {C}^{3\alpha \beta \gamma }&= \mathbb {C}^{\alpha 3 \beta \gamma } = \mathbb {C}^{\alpha \beta 3\gamma }=\mathbb {C}^{\alpha \beta \gamma 3}=0, \\ \mathbb {C}^{333\alpha }&= \mathbb {C}^{3 3 \alpha 3} = \mathbb {C}^{3 \alpha 33}=\mathbb {C}^{\alpha 333}=0,\\ \mathbb {C}^{3333}&= \lambda +2\mu , \end{aligned} \end{aligned}$$
(5)
with \(\lambda\) and \(\mu\) denoting the Lamé constants. For a given displacement field \({\mathbf u}\) the covariant components of the linearized strain \(\pmb {\epsilon }= \epsilon _{ij} \; \mathbf {G}^{i} \otimes \mathbf {G}^{j}\) are
$$\begin{aligned} \begin{aligned} \epsilon _{\alpha \beta }&= \frac{1}{2}\left( \mu _{\alpha }^{\varsigma } \mathbf {\overline{G}}_{\varsigma } \cdot \frac{\partial {\mathbf u}(\theta ^j) }{\partial \theta ^\beta } + \mu _{\beta }^{\varsigma } \mathbf {\overline{G}}_{\varsigma } \cdot \frac{\partial {\mathbf u}(\theta ^j) }{\partial \theta ^\alpha } \right) , \\ \epsilon _{\alpha 3}&= \frac{1}{2}\left( \mu _{\alpha }^{\varsigma } \mathbf {\overline{G}}_{\varsigma } \cdot \frac{\partial {\mathbf u}(\theta ^j) }{\partial \theta ^3} + \mathbf {n}\cdot \frac{\partial {\mathbf u}(\theta ^j) }{\partial \theta ^\alpha } \right) ,\\ \epsilon _{33}&= \frac{\partial {\mathbf u}(\theta ^j) }{\partial \theta ^3} \cdot \mathbf {n}, \end{aligned} \end{aligned}$$
(6)
with the components of the shifter tensor \(\mu _{\alpha }^{\beta } =\left( \delta _{\alpha }^{\beta } - \theta ^3 h_{\alpha }^{\beta } \right)\), where \(h^\beta _\alpha = G^{\beta \gamma }\, \mathbf {G}_{\alpha ,\gamma }\cdot \mathbf {n}\).
One common assumption in shell theories is that strait fibers normal to the mid-surface remain strait after deformation. Therefore, we assume a displacement field of the form
$$\begin{aligned} \bar{{\mathbf u}}(\theta ^1,\theta ^2,\theta ^3) = u^i(\theta ^1,\theta ^2) \mathbf {e}_{i} + \theta ^3 v^\alpha (\theta ^1,\theta ^2) \mathbf {\overline{G}}_{\alpha }. \end{aligned}$$
(7)
Kirchhoff–Love model
In the Kirchhoff–Love model \(v^\alpha (\theta ^1,\theta ^2)\) are no independent parameters. The assumption that normals to the undeformed surface remain normal to the deformed surface and remain unstreched leads to
$$\begin{aligned} v^\alpha = -\overline{G}^{\alpha \gamma } {\mathbf u}_{,\gamma } \cdot \mathbf {n}\end{aligned}$$
(8)
and
$$\begin{aligned} \bar{{\mathbf u}}(\theta ^1,\theta ^2,\theta ^3) = u^i(\theta ^1,\theta ^2) \mathbf {e}_{i} - \theta ^3 \overline{G}^{\alpha \gamma } ({\mathbf u}_{,\gamma } \cdot \mathbf {n}) \mathbf {\overline{G}}_{\alpha }. \end{aligned}$$
(9)
Therefore, the components of the consistent strain tensor (6) can be computed
$$\begin{aligned} \begin{aligned} \epsilon _{11} =\,&\mu _{1}^{\varsigma } u^i_{,1} \overline{J}_{\varsigma i} - \theta ^3 \mu _{1}^{\varsigma } \left( u^i_{,\varsigma 1} \overline{J}_{3 i} - u^i_{,\varsigma }\overline{J}_{\beta i} h_1^\beta - u^i_{,\gamma } \overline{J}_{3 i}\overline{\varGamma }_{1\varsigma }^{\;\;\;\; \gamma } \right) , \\ 2\epsilon _{12} =\,&\mu _{2}^{\varsigma } u^i_{,1} \overline{J}_{\varsigma i}+\mu _{1}^{\varsigma } u^i_{,2} \overline{J}_{\varsigma i} - \theta ^3 \mu _{2}^{\varsigma } \left( u^i_{,\varsigma 1} \overline{J}_{3 i} - u^i_{,\varsigma }\overline{J}_{\beta i} h_1^\beta - u^i_{,\gamma } \overline{J}_{3 i}\overline{\varGamma }_{1\varsigma }^{\;\;\;\; \gamma } \right) \\&-\,\theta ^3 \mu _{1}^{\varsigma } \left( u^i_{,\varsigma 2} \overline{J}_{3 i} - u^i_{,\varsigma }\overline{J}_{\beta i} h_2^\beta - u^i_{,\gamma } \overline{J}_{3 i}\overline{\varGamma }_{2\varsigma }^{\;\;\;\; \gamma } \right) ,\\ \epsilon _{22} =\,&\mu _{2}^{\varsigma } u^i_{,2} \overline{J}_{\varsigma i} - \theta ^3 \mu _{2}^{\varsigma } \left( u^i_{,\varsigma 2} \overline{J}_{3 i} - u^i_{,\varsigma }\overline{J}_{\beta i} h_2^\beta - u^i_{,\gamma } \overline{J}_{3 i}\overline{\varGamma }_{2\varsigma }^{\;\;\;\; \gamma } \right) ,\\ \epsilon _{\alpha 3} =\,&0, \\ \epsilon _{33} =\,&0. \end{aligned} \end{aligned}$$
(10)
In order to avoid Poisson thickness locking the zero transverse normal stress assumption (\(\sigma _{33} = 0\)) has to be included. Enforcing the condition leads to the expression
$$\begin{aligned} \epsilon _{33} = \frac{\lambda }{\lambda +2\mu }\left( \mathrm G^{11}\epsilon _{11}+2\mathrm G^{12}\epsilon _{12} + \mathrm G^{22}\epsilon _{22}\right) \end{aligned}$$
(11)
for the modified transverse normal strain. Within the finite element code this results in a modification in the material law, i.e. in all equations \(\lambda\) has to be replaced with \(\frac{2\mu \lambda }{2\mu +\lambda }\).
Reissner–Mindlin model
In the Reissner–Mindlin model \(v^\alpha (\theta ^1,\theta ^2)\) are two additional independent unknown fields. The components of the strain tensor (6) are now
$$\begin{aligned} \begin{aligned} \epsilon _{11}&= \mu _{1}^{\varsigma } \overline{J}_{\varsigma i} u^i_{\,,1} + \theta ^3 \mu _{1}^{\varsigma } \left( v^\gamma _{\,,1} \overline{G}_{\varsigma \gamma }+ v^\gamma \overline{\varGamma }_{\gamma 1 \varsigma } \right) , \\ 2\epsilon _{12}&= \mu _{1}^{\varsigma } \overline{J}_{\varsigma i} u^i_{\,,2}+\mu _{2}^{\varsigma } \overline{J}_{\varsigma i} u^i_{\,,1}+ \theta ^3 \mu _{1}^{\varsigma } \left( v^\gamma _{\,,2} \overline{G}_{\varsigma \gamma }+ v^\gamma \overline{\varGamma }_{\gamma 2 \varsigma } \right) + \theta ^3 \mu _{2}^{\varsigma } \left( v^\gamma _{\,,1} \overline{G}_{\varsigma \gamma }+ v^\gamma \overline{\varGamma }_{\gamma 1 \varsigma } \right) , \\ \epsilon _{22}&=\mu _{2}^{\varsigma } \overline{J}_{\varsigma i} u^i_{\,,2} + \theta ^3 \mu _{2}^{\varsigma } \left( v^\gamma _{\,,2} \overline{G}_{\varsigma \gamma }+ v^\gamma \overline{\varGamma }_{\gamma 2 \varsigma } \right) ,\\ 2\epsilon _{\alpha 3}&= \mu _{\alpha }^{\varsigma } \overline{G}_{\varsigma \gamma } v^\gamma + u^i_{\,,\alpha } J_{3i} + v^\gamma h_{\gamma \alpha }, \\ \epsilon _{33}&= 0. \end{aligned} \end{aligned}$$
(12)
Again the zero transverse normal stress condition has to be enforced in order to avoid Poisson thickness locking.
Finite element method
In this section, we describe the used finite element approach briefly. We have the following variational formulation of (1):
Find \({\tilde{{\mathbf u}}}\in V\) such that
$$\begin{aligned} \int _{\varOmega } {\tilde{\pmb {\epsilon }}}({\tilde{\mathbf{v}}}) : {\tilde{\mathbb {C}}} : {\tilde{\pmb {\epsilon }}}({\tilde{{\mathbf u}}}) \;{\mathrm{d}}{\mathbf{x}} = \int _{\varOmega } {\tilde{\mathbf{v}}} \cdot {\tilde{{\mathbf b}}} \;{\mathrm {d}}{\mathbf{x}} + \int _{\varGamma _N} {\tilde{\mathbf{v}}} \cdot {\tilde{{\mathbf t}}} \;{\mathrm {d}}s_{\mathbf{x}} \quad \forall \,{\tilde{\mathbf{v}}} \in V_0. \end{aligned}$$
(13)
Here, \(V=\{{\tilde{{\mathbf u}}} \in [H^1(\varOmega )]^3\;|\;{\tilde{{\mathbf u}}}={\tilde{{\mathbf u}}} \;\text {on}\; \varGamma _D \}\) and \(V_0=\{{\tilde{{\mathbf u}}} \in [H^1(\varOmega )]^3\;|\;{\tilde{{\mathbf u}}} = 0 \;\text {on}\; \varGamma _D \}\). We consider the change of variables according to \({\mathbf u}(\theta ^1,\theta ^2,\theta ^3) = {\tilde{{\mathbf u}}}(\mathfrak {g}(\theta ^1,\theta ^2,\theta ^3))\) for all quantities in (13). Thus, the integrals in (13) are evaluated on the parametric domain utilizing a quadrature rule. In particular, we use a tensor product quadrature rule composed of one-dimensional Gauss–Legendre quadratures for the volume integrals. Therefore, we have for a generic integrand \({\mathcal{A}}(\theta ^1,\theta ^2,\theta ^3) = {\mathcal {{\tilde{A}}}}(x)\),
$$\begin{aligned} \begin{aligned} \int _{\varOmega } {\mathcal {{\tilde{A}}}}(x) \;{\mathrm {d}}{\mathbf{x}}&=\int _{{\bar{U}}\times T} {\mathcal{A}}(\theta ^1,\theta ^2,\theta ^3) (1-\theta ^32H+(\theta ^3)^2 K ) \sqrt{\det \bar{G}_{\alpha \beta }} \;{\mathrm {d}}\theta ^1 \;{\mathrm {d}}\theta ^2 \;{\mathrm {d}}\theta ^3 \\&\approx \sum _{i=1}^{n_1} \sum _{j=1}^{n_2}\sum _{k=1}^{n_3} {\mathcal{A}}(\theta ^1_i,\theta ^2_j,\theta ^3_k) (1-\theta ^3_k \, 2H+(\theta ^3_k)^2 K ) \sqrt{\det \bar{G}_{\alpha \beta }}\;\omega _i\omega _j\omega _k, \end{aligned} \end{aligned}$$
(14)
where \(H=\frac {1}{2}(h_\alpha ^\alpha )\) is the mean curvature and \(K=\det (h_\alpha ^\beta )\) is the Gaussian curvature. For \(a=1,2,3\), \(n_a\) is the number of quadrature points \(\theta ^a_i\) and quadrature weights \(\omega _i\).
In order to discretize the sought field \({\mathbf u}\) we apply two steps. First, we resolve the trough-the-thickness variation (the dependency on \(\theta ^3\)) of \({\mathbf u}\) either by the Kirchhoff–Love kinematics (9) or the Reissner–Mindlin kinematics (7). The model decision has great impact on the subsequent discretization done in a second step. In the case of the Reissner–Mindlin shell only first order derivatives appear in (12). Therefore, we can use \(H^1\) hierarchical shape functions on quadrilaterals [23] for the discretization of the five components \(u^i\) and \(v^\alpha\). Due to the second derivatives in (10) this approach is not feasible for the Kirchhoff–Love model. In order to have a \(H^2\) conforming method we use the Bogner–Fox–Schmidt quadrilateral [24] to discretize the three components \(u^i\). Dirichlet boundary conditions are incorporated in the ansatz space in the case of the Reissner–Mindlin model or enforced by a penalty technique in the case of the Kirchhoff–Love model (cf. [25]).