1 Introduction

The relation between solutions of the equations of nonlinear elasticity and solutions for lower-dimensional models is of great interest since the lower-dimensional models are often easier to analyze and to use for numerical simulations. A general introduction to this topic can be found in [5] or for continuum mechanics see [9]. In dependence of the size of the deformation and the applied forces, different lower-dimensional models can occur. Therefore, a rigorous derivation of the lower-dimensional models is of great interest. In the case of the time-independent case, there are many results, cf.  e.g., Friesecke et al. [7, 8] for the case of plates and Mora and Müller [11, 12] in the case of rods and Scardia [14, 15] for curved rods. But there are only few results in the time-independent case so far.

In this contribution we investigate the relation between solutions of an appropriately scaled wave equation of nonlinear elasticity and solutions of a linear Euler–Bernoulli beam system. More precisely, let \(\Omega := [0,L]\times S\) be the reference configuration of a three-dimensional rod, where \(L>0\) and \(S\subset \mathbb {R}^2\) is the cross section. Then we consider the following nonlinear system

$$\begin{aligned}&\partial _t^2 u_h - \frac{1}{h^2}{\text {div}}_h \Big (D\tilde{W}(\nabla _h u_h)\Big ) = h^{2} f_h \quad \text {in } \Omega \times [0, T), \end{aligned}$$
(1.1)
$$\begin{aligned}&D\tilde{W}(\nabla _h u_h)\nu |_{(0, L)\times \partial S} = 0, \end{aligned}$$
(1.2)
$$\begin{aligned}&u_h \text { is} \,L \,\text {-periodic} \text { w.r.t. } x_1, \end{aligned}$$
(1.3)
$$\begin{aligned}&(u_h, \partial _t u_h)|_{t=0} = (u_{0,h}, u_{1,h}), \end{aligned}$$
(1.4)

where \(T > 0\) and \(\tilde{W}\) is some elastic energy density chosen later. Existence of strong solutions for large times of this system was shown in [1]. More details on how to justify the scaling can be found there as well. The limit system as \(h\rightarrow 0\) is given by

$$\begin{aligned}{} & {} \partial _t^2 v + \begin{pmatrix} I_2 &{} 0 \\ 0 &{} I_3 \end{pmatrix} \partial ^4_{x_1} v = g\quad \text {in } [0,L]\times [0,\infty ) \end{aligned}$$
(1.5)
$$\begin{aligned}{} & {} v \text { is }\,L\,\text {-periodic in}\,x_1 \end{aligned}$$
(1.6)
$$\begin{aligned}{} & {} (v,\partial _t v)|_{t=0} = (\tilde{v}_0, \tilde{v}_1) \end{aligned}$$
(1.7)

where \(\tilde{v}_0, \tilde{v}_1\) and \(I_2, I_3\) are appropriately chosen initial values and weights, respectively. It is the goal of this manuscript to prove convergence of the solutions of the system (1.1)–(1.4) to solutions of the limit system (1.5)–(1.7) with appropriate convergence rates for well-prepared initial data, cf. Theorem 3.3 below.

In an energetic setting the relations between higher-dimensional models and lower-dimensional ones, using the notion of \(\Gamma \)-convergence a fundamental contribution was given in [7]. There the classical geometric rigidity is proven. Using this result it was possible to prove a lot of convergence results in different geometrical situations and scaling regimes in the static setting, see, for instance, [8, 11, 12]. In the dynamical case for plates, a convergence result can be found in [3] and in [6] in the case of viscoelasticity. The large times existence and a first-order asymptotic for plates were shown in [2].

In the following we want to explain the main novelties and difficulties of this contribution. In a first step we construct an approximation using the solution of the lower-dimensional system. This approximation is constructed such that it solves the linearization around zero of the nonlinear, three-dimensional equation up to an error of order \(h^3\). This is done explicitly by determining suitable prefactor functions as solutions of systems on S. Thereafter, the main difficulty of this work is to establish existence of suitable initial data in order to ensure large times existence for the solution of the nonlinear problem. This is done in Sect. 3.2. Here we use the nonlinear equations for the initial data from the compatibility conditions. These are solved via a fixed point argument on precisely chosen function spaces. Finally, in Sect. 3.3 the convergence properties are proven. For this we use a general result for solutions of the linearized equation. Moreover, we have to carefully treat the rotational parts of the initial data, as the spaces for the fixed point argument do not cover them. For this we use a decomposition and the fact that the elastic energy density is chosen as \(\tilde{W}(F) = {\text {dist}}(Id+F; SO(3))\).

The results are part of the second author’s PhD thesis [4].

2 Preliminaries and auxiliary results

2.1 Notation

We use standard notation; in particular \(\mathbb {N}\) and \(\mathbb {N}_0 := \mathbb {N}\cup \{0\}\) denote the natural numbers without and with zero, respectively. Moreover, the norm on \(\mathbb {R}\) and absolute value in \(\mathbb {R}^n\), \(\mathbb {R}^{n\times n}\) is denoted by |.| for all \(n\in \mathbb {N}\). For \(p,~k\in \mathbb {N}\), we denote the classical Lebesgue and Sobolev spaces for some bounded, open set \(M\subset \mathbb {R}^n\), by \(L^p(M)\), \(W^k_p(M)\) and \(H^k(M) := W^k_2(M)\). A subscript (0) on a function space will always indicate that elements have zero mean value, e.g., for \(g\in H^1_{(0)}(M)\) we have

$$\begin{aligned} \int \limits _M g(x) \textrm{d}x = 0. \end{aligned}$$
(2.1)

The cross section of the rod is always denoted by \(S\subset \mathbb {R}^2\) and is assumed to be a smooth and bounded domain. Furthermore, let \(\Omega _h := (0,L) \times hS \subset \mathbb {R}^3\) for \(h \in (0,1]\) and \(L>0\) and for convenience we write \(\Omega := \Omega _1\). We assume that S satisfies

$$\begin{aligned}&\int \limits _S x_2 x_3 \textrm{d}x' = 0\quad \text {and}\end{aligned}$$
(2.2)
$$\begin{aligned}&\int \limits _S x_2 \textrm{d}x' = \int \limits _S x_3 \textrm{d}x' = 0, \end{aligned}$$
(2.3)

where \(x' := (x_2, x_3)\subset \mathbb {R}^2\). This is no loss of generality, as it can always be achieved via a translation and rotation. The scaling shell is such that we can assume \(|S|=1\). Furthermore, we denote with \(\nabla _h\) the scaled gradient defined as

$$\begin{aligned} \nabla _h = \bigg (\partial _{x_1}, \frac{1}{h}\partial _{x_2}, \frac{1}{h}\partial _{x_3}\bigg )^T\quad \text {and}\quad \varepsilon _h(u)= {\text {sym}}(\nabla _h u). \end{aligned}$$

The respective gradient in only \(x':=(x_2, x_3)\) direction is denoted by

$$\begin{aligned} \nabla _{x'} := \big (\partial _{x_2}, \partial _{x_3}\big )^T. \end{aligned}$$

The standard notation \(H^k(\Omega )\) and \(H^k(\Omega ; X)\) is used for \(L^2\)-Sobolev spaces of order \(k\in \mathbb {N}\) with values in \(\mathbb {R}\) and some space X, respectively.

The space of all n-linear mappings \(G:V^n \rightarrow \mathbb {R}\) for a vector space V is denoted, throughout the paper by \(\mathcal {L}^n(V)\), \(n\in \mathbb {N}\). We deploy the standard identification of \(\mathcal {L}^1(\mathbb {R}^{n\times n}) = (\mathbb {R}^{n\times n})'\) with \(\mathbb {R}^{n\times n}\), i.e., \(G\in \mathcal {L}^1(\mathbb {R}^{n\times n})\) is identified with \(A\in \mathbb {R}^{n\times n}\) such that

$$\begin{aligned} G(X) = A : X \quad \text {for all}\quad X\in \mathbb {R}^{n\times n}\end{aligned}$$

where \(A:X = \sum _{i,j=1}^n a_{ij} x_{ij}\) is the usual inner product on \(\mathbb {R}^{n\times n}\). Analogously, for \(G\in \mathcal {L}^2(\mathbb {R}^{n\times n})\) we use the identification with \(\tilde{G}:\mathbb {R}^{n\times n}\rightarrow \mathbb {R}^{n\times n}\) defined by

$$\begin{aligned} \tilde{G}X : Y = G(X,Y)\quad \text {for all}\quad X, Y\in \mathbb {R}^{n\times n}. \end{aligned}$$
(2.4)

We introduce a scaled inner product on \(\mathbb {R}^{n\times n}\)

$$\begin{aligned} A:_hB := \frac{1}{h^2} {\text {sym}}A : {\text {sym}}B + {\text {skew}}A : {\text {skew}}B \end{aligned}$$

for all A, \(B\in \mathbb {R}^{n\times n}\) and \(h>0\) and the corresponding norm is denoted by \(|A|_h := \sqrt{A:_hA}\). With this we can define for \(W\in \mathcal {L}^d(\mathbb {R}^{n\times n})\) the induced scaled norm by

$$\begin{aligned} |W|_h := \sup _{|A_j|_h \le 1, j =\{1,\ldots , d\}} |W(A_1, \ldots , A_d)| \end{aligned}$$

Using \(|A|_h \ge |A|_1 = |A|\) for all \(A\in \mathbb {R}^{n\times n}\), it follows \(|W|_h \le |W|_1 =: |W|\) for all \(W\in \mathcal {L}^d(\mathbb {R}^{n\times n})\) and \(0 < h \le 1\). The scaled \(L^p\)-spaces are defined as follows

$$\begin{aligned} \Vert W\Vert _{L^p_h(U, \mathcal {L}^d(\mathbb {R}^{n\times n}))} = \Vert W\Vert _{L^p_h(U)} = \left( \int \limits _U |W(x)|_h^p dx\right) ^{\frac{1}{p}} \end{aligned}$$

if \(p\in [1, \infty )\), where \(U\subset \mathbb {R}^d\) is measurable. Thus \(\Vert W\Vert _{L^p_h(U; \mathcal {L}^d(\mathbb {R}^{n\times n}))} \le \Vert W\Vert _{L^p(U;\mathcal {L}^d(\mathbb {R}^{n\times n}))}\). The scaled norm for \(f\in L^p(U, \mathbb {R}^{n\times n})\) is defined in the same way

$$\begin{aligned} \Vert f\Vert _{L^p_h(U,\mathbb {R}^{n\times n})} = \Vert f\Vert _{L^p_h(U)} = \left( \int \limits _U |f(x)|_h^p dx\right) ^{\frac{1}{p}}. \end{aligned}$$

Then

$$\begin{aligned} \Vert f\Vert _{L^p_h(U; \mathbb {R}^{n\times n})} \ge \Vert f\Vert _{L^p(U;\mathbb {R}^{n\times n})}. \end{aligned}$$

As we will work with periodic boundary condition in \(x_1\)-direction, we introduce for \(m\in \mathbb {N}\)

$$\begin{aligned} H^m_{\textrm{per}}(\Omega ) := \Big \{f\in H^m(\Omega ) \;:\; \partial ^\alpha _x f|_{x_1 = 0} = \partial ^\alpha _x f|_{x_1 = L},\;\text {for all } |\alpha | \le m-1 \Big \}. \end{aligned}$$

This space can equivalently defined in the following way, which is in some situations more convenient

$$\begin{aligned} \tilde{H}^m_{\textrm{per}}(\Omega ) := \Big \{f\in H^m_{loc}(\mathbb {R}\times {\bar{S}}):\,f(x_1, x')= f(x_1 + L, x') \text { almost everywhere}\Big \} \end{aligned}$$

We equipped \(\tilde{H}^m_{\textrm{per}}(\Omega )\) with the standard \(H^m(\Omega )\)-norm. As the maps \(f\mapsto f|_{\Omega }:\tilde{H}^m_{\textrm{per}}(\Omega ) \rightarrow H^m_{\textrm{per}}(\Omega )\) and \(f\mapsto f_{\textrm{per}}:H^m_{\textrm{per}}(\Omega ) \rightarrow \tilde{H}^m_{\textrm{per}}(\Omega )\) are isomorphisms, we identify \(\tilde{H}^m_{\textrm{per}}(\Omega )\) with \(H^m_{\textrm{per}}(\Omega )\). This leads immediately to the density of smooth functions in \(H^m_{\textrm{per}}(\Omega )\), because, as S is smooth, there exists an appropriate extension operator and thus we can use a convolution argument.

In various estimates we will use an anisotropic variant of \(H^k(\Omega )\), as we will have more regularity in lateral direction. Therefore, we define

$$\begin{aligned} H^{m_1, m_2} (\Omega )&:= \Big \{u\in L^2(\Omega ) \;:\; \partial _{x_1}^l \nabla _{x}^k u\in L^2(\Omega ), \text { for } k=0,\ldots , m_1, l=0,\ldots ,m_2\\&\quad \partial _{x_1}^q \partial _{x}^\alpha u\Big |_{x_1 = 0} = \partial _{x_1}^q\partial _{x}^\alpha u\Big |_{x_1 = L}, \text { for } q = 0,\ldots , m_1, |\alpha |\le m_2 \text { and } q + |\alpha | \le m_1 + m_2 -1 \Big \} \end{aligned}$$

where \(m_1,~m_2\in \mathbb {N}_0\), the inner product is given by

$$\begin{aligned} (f,g)_{H^{m_1, m_2}(\Omega )} = \sum _{k=0,\ldots ,m_1; l=0,\ldots m_2} \Big (\partial _{x_1}^l \nabla _x^k f, \partial _{x_1}^l \nabla _x^k g \Big )_{L^2(\Omega )}. \end{aligned}$$

Furthermore, we will use the scaled norms

$$\begin{aligned} \Vert A\Vert _{H^m_h(\Omega )}&:= \left( \sum _{|\alpha | \le m} \Vert \partial ^\alpha _x A\Vert _{L^2_h(\Omega )}^2\right) ^{\frac{1}{2}}\\ \Vert B\Vert _{H^{m_1, m_2}_h(\Omega )}&:= \left( \sum _{k=0,\ldots ,m_1; l=0,\ldots ,m_2} \Vert \partial _{x_1}^l \nabla _x^k B\Vert ^2_{L^2_h(\Omega )}\right) ^{\frac{1}{2}}. \end{aligned}$$

for \(A\in H^m(\Omega ;\mathbb {R}^{n\times n})\) and \(B\in H^{m_1,m_2}(\Omega ;\mathbb {R}^{n\times n})\) and \(n\in \mathbb {N}\). As an abbreviation we denote for \(u\in H^k(\Omega ;\mathbb {R}^3)\) the symmetric scaled gradient by \(\varepsilon _h(u):={\text {sym}}(\nabla _h u)\) and \(\varepsilon (u) = \varepsilon _1(u) = {\text {sym}}(\nabla u)\).

The following lemma provides the possibility to take traces for \(u\in H^{0,1}(\Omega )\):

Lemma 2.1

The operator \({\text {tr}}_{a}:H^{0,1}(\Omega ) \rightarrow L^2(S)\), \(u\mapsto u|_{x_1=a}\) is well defined and bounded.

Proof

This is an immediate consequence of the embedding

$$\begin{aligned} H^{0,1}(\Omega ) = H^1(0,L;L^2(S)) \hookrightarrow BUC([0,L];L^2(S)) \end{aligned}$$

where BUC([0, L]; X) is the space of all uniformly continuous functions \(f:[0,L] \rightarrow X\) for some Banach space X. \(\square \)

2.2 The strain energy density W and Korn’s inequality

We investigate the mathematical assumptions and resulting properties of the strain-energy density W we use in this contribution. We assume to have \(W:\mathbb {R}^{3\times 3}\rightarrow [0,\infty )\) defined by

$$\begin{aligned} W(F) := \frac{1}{2} {\text {dist}}(F, SO(3)) \end{aligned}$$

where SO(3) denotes the group of special orthogonal matrices. This energy density clearly satisfies the following general assumptions

  1. (i)

    \(W\in C^\infty (B_\delta (Id); [0,\infty ))\) for some \(\delta >0\);

  2. (ii)

    W is frame-invariant, i.e., \(W(RF) = W(F)\) for all \(F\in \mathbb {R}^{3\times 3}\) and \(R\in SO(3)\);

  3. (iii)

    there exists \(c_0 > 0\) such that \(W(F) \ge c_0 {\text {dist}}(F, SO(3))^2\) for all \(F\in \mathbb {R}^{3\times 3}\) and \(W(R) = 0\) for every \(R\in SO(3)\).

Remark 2.2

We note that W has a minimum point at the identity, as \(W(Id) = 0\) and \(W(F)\ge 0\) for all \(F\in \mathbb {R}^{3\times 3}\). Hence, we have for \(\tilde{W}(F):=W(Id + F)\) for all \(F\in \mathbb {R}^{3\times 3}\), \(D\tilde{W}(0)[G] = 0\) for all \(G\in \mathbb {R}^{3\times 3}\). Moreover, it holds \(D^2 \tilde{W}(0) F = {\text {sym}}F\) and for \(P\in \mathbb {R}^{3\times 3}_{skew}\), \(A, B\in \mathbb {R}^{3\times 3}\) we obtain

$$\begin{aligned} D^3 \tilde{W}(0)[A,B,P] = \Big ((A^T- A)^T {\text {sym}}(B) + (B^T - B)^T {\text {sym}}(A)\Big ) : P. \end{aligned}$$
(2.5)

The following lemma provides an essential decomposition of \(D^3\tilde{W}\) in the general form.

Lemma 2.3

There is some constant \(C>0\), \(\varepsilon >0\) and \(A\in C^\infty (\overline{B_\varepsilon (0)}; \mathcal {L}^3(\mathbb {R}^{n\times n}))\) such that for all \(G\in \mathbb {R}^{n\times n}\) with \(|G| \le \varepsilon \) we have

$$\begin{aligned} D^3\tilde{W}(G) = D^3\tilde{W}(0) + A(G) \end{aligned}$$

where

$$\begin{aligned}&|D^3\tilde{W}(0)|_h \le Ch\quad \text {for all}\quad 0<h\le 1, \end{aligned}$$
(2.6)
$$\begin{aligned}&|A(G)| \le C|G|\quad \text {for all}\quad |G|\le \varepsilon . \end{aligned}$$
(2.7)

Proof

For the proof we refer to [2, Lemma 2.6]. \(\square \)

With this we can prove the following bound for \(D^3\tilde{W}\).

Corollary 2.4

There exist C, \(\varepsilon > 0\) such that

$$\begin{aligned} \Vert D^3\tilde{W}(Z)(Y_1, Y_2, Y_3)\Vert _{L^1(\Omega )} \le Ch \Vert Y_1\Vert _{H^2_h(\Omega )}\Vert Y_2\Vert _{L^2_h(\Omega )}\Vert Y_3\Vert _{L^2_h(\Omega )} \end{aligned}$$
(2.8)

for all \(Y_1\in H^2(\Omega , \mathbb {R}^{n\times n})\), \(Y_2\), \(Y_3\in L^2(\Omega ; \mathbb {R}^{n\times n})\), \(0<h\le 1\) and \(\Vert Z\Vert _{L^\infty (\Omega } \le \min \{\varepsilon , h\}\) and

$$\begin{aligned} \Vert D^3\tilde{W}(Z)(Y_1, Y_2, Y_3)\Vert _{L^1(\Omega )} \le Ch \Vert Y_1\Vert _{H^1_h(\Omega )} \Vert Y_2\Vert _{H^1_h(\Omega )} \Vert Y_3\Vert _{L^2_h(\Omega )} \end{aligned}$$
(2.9)

for all \(Y_1\), \(Y_2\in H^1(\Omega , \mathbb {R}^{n\times n})\), \(Y_3\in L^2(\Omega ; \mathbb {R}^{n\times n})\), \(0<h\le 1\) and \(\Vert Z\Vert _{L^\infty (\Omega } \le \min \{\varepsilon , h\}\) and

$$\begin{aligned} \Vert D^3\tilde{W}(Z)(Y_1, Y_2, Y_3)\Vert _{L^1(\Omega )} \le Ch \bigg \Vert \bigg (Y_1, \frac{1}{h} {\text {sym}}(Y_1)\bigg )\bigg \Vert _{L^\infty (\Omega )} \Vert Y_2\Vert _{H^1_h(\Omega )} \Vert Y_3\Vert _{L^2_h(\Omega )} \end{aligned}$$
(2.10)

for all \(Y_1\in L^\infty (\Omega , \mathbb {R}^{n\times n})\), \(Y_2\), \(Y_3\in L^2(\Omega ; \mathbb {R}^{n\times n})\), \(0<h\le 1\) and \(\Vert Z\Vert _{L^\infty (\Omega } \le \min \{\varepsilon , h\}\).

Proof

The inequalities follow directly from Lemma 2.3 and Hölder’s inequality. \(\square \)

In order to bound the full scaled gradient \(\nabla _h g\) of some function \(g\in H^1_{\textrm{per}}(\Omega )\) by the symmetric one, we need a sharp Korn’s inequality for thin rods. As rigid motions \(x\mapsto \alpha x^\perp \) for \(\alpha \in \mathbb {R}\) arbitrary are admissible functions in \(H^1_{\textrm{per}}(\Omega )\), we cannot expect that the full scaled gradient is bounded by \(\varepsilon _h(g)\). Precisely, we obtain the following results.

Lemma 2.5

There exists a constant \(C=C(\Omega )>0\) such that for all \(0 < h\le 1\) and \(u\in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) we have

$$\begin{aligned} \bigg \Vert \nabla _h u - \frac{1}{h}B(u)\bigg \Vert _{L^2(\Omega )} \le C\bigg \Vert \frac{1}{h} \varepsilon _h(u)\bigg \Vert _{L^2(\Omega )}, \end{aligned}$$
(2.11)

where

$$\begin{aligned} B(u) = \begin{pmatrix} 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad a(u)\\ 0&{}\quad - a(u)&{}\quad 0 \end{pmatrix} \end{aligned}$$
(2.12)

with \(a(u) = \frac{1}{|\Omega |}\int \limits _\Omega \partial _{x_3} u_2(x) - \partial _{x_2} u_3(x) \textrm{d}x\).

Proof

The proof is similar to [2, Lemma 2.1] and is done in [4, Lemma 2.4.4] \(\square \)

Lemma 2.6

(Korn inequality in integral form) For all \(0 < h\le 1\) and \(u\in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\), there exists a constant \(C_K=C_K(\Omega )\), such that

$$\begin{aligned} \Vert \nabla _h u\Vert _{L^2(\Omega )}\le \frac{C_K}{h}\bigg ( \Vert \varepsilon _h(u)\Vert _{L^2(\Omega )} + \bigg | \int \limits _\Omega u \cdot x^\perp \textrm{d}x\bigg |\bigg ) \end{aligned}$$
(2.13)

where \(x^\perp = (0, -x_3, x_2)^T\).

Proof

A proof can be found in [1]. \(\square \)

3 First-order expansion in a Linearized regime

We construct an approximation to the unique solution of the nonlinear system

$$\begin{aligned}&\partial _t^2 u_h - \frac{1}{h^2}{\text {div}}_h \Big (D\tilde{W}(\nabla _h u_h)\Big ) = h^{2} f_h \quad \text {in } \Omega \times [0, T), \end{aligned}$$
(3.1)
$$\begin{aligned}&D\tilde{W}(\nabla _h u_h)\nu |_{(0, L)\times \partial S} = 0, \end{aligned}$$
(3.2)
$$\begin{aligned}&u_h \text { is }\,L\,\text {-periodic} \text { w.r.t. } x_1, \end{aligned}$$
(3.3)
$$\begin{aligned}&(u_h, \partial _t u_h)|_{t=0} = (u_{0,h}, u_{1,h}), \end{aligned}$$
(3.4)

where \(\tilde{W}(F) = W(Id+F)\) for all \(F\in \mathbb {R}^{3\times 3}\), \(T > 0\). We assume that

$$\begin{aligned} f^h(x,t) = \begin{pmatrix} 0 \\ g(x_1, t) \end{pmatrix} \end{aligned}$$

for some \(g\in \bigcap _{k=0}^3 W^k_1(0,T; H^{10-2k}_{\textrm{per}}(0,L; \mathbb {R}^2))\), which implies

$$\begin{aligned} \int \limits _S f^h(x,t) x_k dx' = 0 \end{aligned}$$

for \(k=2, 3\). Moreover, we assume that

$$\begin{aligned} \max _{\sigma = 0,1,2}\Vert \partial _t^\sigma g|_{t=0}\Vert _{H^{2-2\sigma }(0,L)} \le M, \end{aligned}$$
(3.5)

where \(M>0\) is chosen later. Without loss of generality we can assume \(\int \limits _0^L g dx_1 = 0\). Otherwise, we subtract

$$\begin{aligned} a(t) := \frac{1}{|\Omega |}\left( \int \limits _\Omega u_{0,h} \textrm{d}x - t\int \limits _\Omega u_{1,h} \, \textrm{d}x - \int \limits _0^t (t-s) \int \limits _\Omega f^h(s)\, \textrm{d}x\, \textrm{d}s\right) \end{aligned}$$

from \(u_h\) analogously as in the proof of [1, Theorem 3.1].

3.1 Construction of the ansatz function

For the ansatz function we consider the following system of one-dimensional beam equations

$$\begin{aligned}{} & {} \partial _t^2 v + \begin{pmatrix} I_2 &{}\quad 0 \\ 0 &{}\quad I_3 \end{pmatrix} \partial ^4_{x_1} v = g,\\{} & {} v\,\text {is}\,L\text {-periodic in}\,x_1,\\{} & {} (v,\partial _t v)|_{t=0} = (\tilde{v}_0, \tilde{v}_1), \end{aligned}$$

where \(\tilde{v}_{0}\in H^{12}_{\textrm{per}}(0,L;\mathbb {R}^2)\), \(\tilde{v}_{1}\in H^{10}_{\textrm{per}}(0,L;\mathbb {R}^2)\) such that

$$\begin{aligned} \Vert \tilde{v}_0\Vert _{H^8(0,L)} \le M \quad \text { and }\quad \Vert \tilde{v}_1\Vert _{H^5(0,L)} \le M \end{aligned}$$
(3.6)

and

$$\begin{aligned} I_k := \int \limits _S x_k^2 \textrm{d}x'\quad \text {for}\quad k=2,3. \end{aligned}$$

Then we obtain with standard methods, as, e.g., in [13, Theorem 11.8], the existence of a unique solution

$$\begin{aligned} v\in \bigcap _{j=0}^4 C^j([0,T]; H^{12-2j}_{\textrm{per}}(0,L;\mathbb {R}^2)). \end{aligned}$$

Moreover, due to the assumptions for g and the periodicity of v it follows

$$\begin{aligned} \partial _t^2 \int \limits _0^L v dx_1 = 0. \end{aligned}$$

Now we define

$$\begin{aligned} \tilde{u}_h(x,t)&= h^2 \begin{pmatrix} 0 \\ v_2 \\ v_3 \end{pmatrix} + h^3 \begin{pmatrix} -x_2\partial _{x_1} v_{2} - x_3 \partial _{x_1} v_{3} \\ 0 \\ 0 \end{pmatrix} + h^5 \begin{pmatrix} a_2(x') \partial _{x_1}^3 v_2 + a_3(x') \partial _{x_1}^3 v_3 \\ 0 \\ 0 \end{pmatrix}\nonumber \\&\quad + h^6 \begin{pmatrix} 0 \\ b_2(x') \partial _{x_1}^4 v_2 + c_3(x') \partial _{x_1}^4 v_3 \\ b_3(x') \partial _{x_1}^4 v_3 + c_2(x') \partial _{x_1}^4 v_2 \end{pmatrix}, \end{aligned}$$
(3.7)

where a, b, \(c:S\rightarrow \mathbb {R}^2\) are chosen later. Then

$$\begin{aligned} \nabla _h \tilde{u}_h(x,t)&= h^2 \begin{pmatrix} 0 &{}\quad -\partial _{x_1} v_2 &{}\quad -\partial _{x_1} v_3\\ \partial _{x_1} v_2 &{}\quad 0 &{}\quad 0\\ \partial _{x_1} v_3 &{}\quad 0 &{}\quad 0\\ \end{pmatrix} + h^3 \begin{pmatrix} -x_2 \partial _{x_1}^2 v_2 - x_3 \partial _{x_1}^2 v_3 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \end{pmatrix}\\&\quad + h^4 \begin{pmatrix} 0 &{}\quad \partial _{x_2} a_2 \partial _{x_1}^3 v_2 + \partial _{x_2} a_3 \partial _{x_1}^3 v_3 &{}\quad \partial _{x_3} a_2 \partial _{x_1}^3 v_2 + \partial _{x_3} a_3\partial _{x_1}^3 v_3 \\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\end{pmatrix} \\&\quad +h^5 \begin{pmatrix} a_2 \partial _{x_1}^4 v_2 + a_3 \partial _{x_1}^4 v_3 &{}\quad 0 &{}\quad 0\\ 0&{}\quad \partial _{x_2} b_2 \partial _{x_1}^4 v_2 + \partial _{x_2} c_3 \partial _{x_1}^4 v_3 &{}\quad \partial _{x_3} b_2 \partial _{x_1}^4 v_2 +\partial _{x_3} c_3 \partial _{x_1}^4 v_3\\ 0 &{}\quad \partial _{x_2} b_3 \partial _{x_1}^4 v_3 +\partial _{x_2} c_2 \partial _{x_1}^4 v_2 &{}\quad \partial _{x_3} b_3 \partial _{x_1}^4 v_3 + \partial _{x_3} c_2 \partial _{x_1}^4 v_2 \end{pmatrix}\\&\quad + h^6 \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ b_2 \partial _{x_1}^5 v_2 + c_3 \partial _{x_1}^5 v_3 &{}\quad 0 &{}\quad 0 \\ b_3 \partial _{x_1}^5 v_3 + c_2 \partial _{x_1}^5 v_2 &{}\quad 0 &{}\quad 0 \end{pmatrix}. \end{aligned}$$

Thus, with \(D^2 W(Id) F = {\text {sym}} F\) we can derive

$$\begin{aligned} \frac{1}{h^2} {\text {div}}_h(D^2 W(Id) \nabla _h \tilde{u}_h)&= h \begin{pmatrix} \big (\frac{1}{2} \Delta a - (x_2, x_3)^T\big )\cdot \partial _{x_1}^3 v \\ 0 \\ 0 \end{pmatrix}\\&\quad + h^2 \begin{pmatrix} 0 \\ \nabla _{x'} a(x')^T \partial _{x_1}^4 v +\begin{pmatrix} \partial _{x_2}^2 b_2 &{}\quad \partial _{x_2}^2 c_3\\ \partial _{x_3}^2 c_2 &{}\quad \partial _{x_3}^2 b_3 \end{pmatrix} \partial _{x_1}^4 v \end{pmatrix}\\&\quad + \frac{h^2}{2} \begin{pmatrix} 0 \\ \begin{pmatrix} \partial _{x_3} \partial _{x_2} c_2 + \partial _{x_3}^2 b_2 &{}\quad \partial _{x_2}\partial _{x_3} b_3 + \partial _{x_3}^2 c_3 \\ \partial _{x_2}^2 c_2 + \partial _{x_2}\partial _{x_3} b_2 &{}\quad \partial _{x_2}^2 b_3 + \partial _{x_2}\partial _{x_3} c_3 \end{pmatrix} \partial _{x_1}^4 v \end{pmatrix} + r_h(x,t) \end{aligned}$$

for

$$\begin{aligned} r_h(x,t) = O(h^3). \end{aligned}$$

Moreover, for the boundary condition it holds

$$\begin{aligned}&D^2 W(Id)[\nabla _h \tilde{u}_h] \nu = h^4 \begin{pmatrix} \frac{1}{2}(\nabla _{x'} a \nu _{\partial S}) \cdot \partial _{x_1}^3 v \\ 0 \\ 0 \end{pmatrix}\\&\qquad + h^5 \begin{pmatrix} 0 \\ \big (\partial _{x_2} b_2 \nu _2 + \frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2)\nu _3\big ) \partial _{x_1}^4 v_2 + \big (\partial _{x_2} c_3 \nu _2 + \frac{1}{2}(\partial _{x_2} b_3 + \partial _{x_3} c_3) \nu _3) \partial _{x_1}^4 v_3\big )\\ \big (\frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2) \nu _2 + \partial _{x_3} c_2 \nu _3\big ) \partial _{x_1}^4 v_2 + \Big (\frac{1}{2}(\partial _{x_2} b_3 + \partial _{x_3} c_3)\nu _2 + \partial _{x_3} b_3 \nu _3\Big ) \partial _{x_1}^4 v_3 \end{pmatrix}\\&\qquad + \frac{h^6}{2} \begin{pmatrix} \nu ^T \begin{pmatrix} b_2 &{} c_2 \\ b_3 &{} c_3 \end{pmatrix} \partial _{x_1}^5 v \\ 0 \\ 0 \end{pmatrix}\\&\quad = h^4 \begin{pmatrix} (\nabla _{x'} a \nu _{\partial S}) \cdot \partial _{x_1}^3 v \\ 0 \\ 0 \end{pmatrix} + h^5 \begin{pmatrix} 0 \\ \nu ^T \begin{pmatrix} \partial _{x_2} b_2 &{} \frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2) \\ \partial _{x_2} c_3 &{} \frac{1}{2} (\partial _{x_2} b_3 + \partial _{x_3} c_3) \end{pmatrix} \partial _{x_1}^4 v \\ \nu ^T \begin{pmatrix} \frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2) &{} \partial _{x_3} c_2 \\ \frac{1}{2}(\partial _{x_2} b_3 + \partial _{x_3} c_3) &{} \partial _{x_3} b_3 \end{pmatrix}\partial _{x_1}^4 v \end{pmatrix}\\&\qquad + \frac{h^6}{2} \begin{pmatrix} \nu ^T \begin{pmatrix} b_2 &{} c_2 \\ b_3 &{} c_3 \end{pmatrix} \partial _{x_1}^5 v \\ 0 \\ 0 \end{pmatrix}. \end{aligned}$$

We choose now \(a:S\rightarrow \mathbb {R}^2\) as the solution of the following system

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta a = -2\begin{pmatrix} x_2 \\ x_3 \end{pmatrix} &{}\quad \text {in}\quad S\\ \nabla _{x'} a \nu = 0&{}\quad \text {on}\quad \partial S \end{array}\right. \end{aligned}$$

with

$$\begin{aligned} \int \limits _S a(x') dx' = 0. \end{aligned}$$

Such a solution exists, because we can apply the Lax–Milgram Lemma for the weak Laplacian on \(H^1_{(0)}(S;\mathbb {R}^2)\). Thereby, the coercivity follows from Poincaré’s inequality. With well-known regularity result, e.g., Theorem 4.18 in [10], we obtain \(a\in C^\infty ({\overline{S}},\mathbb {R}^2)\). The systems for b and c decouple to

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{x_2}^2 b_2 + \frac{1}{2} \partial _{x_3}^2 b_2 + \frac{1}{2}\partial _{x_3}\partial _{x_2} c_2 = I_1 - \partial _{x_2} a_2 &{}\quad \text {in}\quad S\\ \frac{1}{2} \partial _{x_2}^2 c_2 + \partial _{x_3}^2 c_2 + \frac{1}{2} \partial _{x_2} \partial _{x_3} b_2 = -\partial _{x_3} a_2 &{}\quad \text {in}\quad S \end{array}\right. \end{aligned}$$
(3.8)

and

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{x_2}^2 c_3 + \frac{1}{2} \partial _{x_3}^2 c_3 + \frac{1}{2}\partial _{x_2}\partial _{x_3} b_3 = - \partial _{x_2} a_3 &{}\quad \text {in}\quad S\\ \frac{1}{2} \partial _{x_2}^2 b_3 + \partial _{x_3}^2 b_3 + \frac{1}{2} \partial _{x_2} \partial _{x_3} c_3 = I_2 -\partial _{x_3} a_3 &{}\quad \text {in}\quad S \end{array}\right. \end{aligned}$$
(3.9)

Defining the matrix of coefficients \(({\mathfrak {p}}^{\alpha \beta }_{ij})^{\alpha , \beta = 2,3}_{i,j=1,2}\) in the following way

$$\begin{aligned} {\mathfrak {p}}^{22}_{11}= & {} 1\quad {\mathfrak {p}}^{33}_{11} = \frac{1}{2}\quad {\mathfrak {p}}^{32}_{12} = \frac{1}{4} \quad {\mathfrak {p}}^{23}_{12} = \frac{1}{4}\\ {\mathfrak {p}}^{22}_{22}= & {} \frac{1}{2}\quad {\mathfrak {p}}^{33}_{22} = 1\quad {\mathfrak {p}}^{23}_{21} = \frac{1}{4} \quad {\mathfrak {p}}^{32}_{21} = \frac{1}{4}\\ {\mathfrak {p}}^{\alpha \beta }_{ij}= & {} 0\quad \text {otherwise}. \end{aligned}$$

With \(w = (b_2, c_2)^T\) and \(f = (-I_1 - \partial _{x_2} a_2, -\partial _{x_3} a_2)^T\), (3.8) is equivalent to

$$\begin{aligned} \sum _{\alpha ,\beta =2}^3 \sum _{j=1}^2 -\partial _\beta \big ({\mathfrak {p}}^{\alpha \beta }_{ij} \partial _\alpha w_j\big ) = f_i \end{aligned}$$

for \(i=1\), 2. Let now

$$\begin{aligned} \xi := \begin{pmatrix} \xi _{12} &{} \xi _{13} \\ \xi _{22} &{} \xi _{23} \end{pmatrix}\in \mathbb {R}^{2\times 2}. \end{aligned}$$

be arbitrary. Then it holds

$$\begin{aligned} \sum _{\alpha ,\beta =2}^3 \sum _{i,j=1}^2 {\mathfrak {p}}^{\alpha \beta }_{ij} \xi _{i\alpha } \xi _{j\beta }&= \frac{3}{4}\left( \xi _{12}^2 + \xi _{23}^2\right) + \frac{1}{4} (\xi _{12} + \xi _{23})^2 + \frac{1}{4}\left( \xi _{13}^2 + \xi _{22}^2\right) + \frac{1}{4} (\xi _{13} + \xi _{22})^2 \\&\ge \frac{1}{4} \left( \xi _{12}^2 + \xi _{13}^2 + \xi _{22}^2 + \xi _{23}^2\right) = \frac{1}{4} |\xi |^2 \end{aligned}$$

and thus \({\mathfrak {p}}^{\alpha \beta }_{ij}\) satisfies the Legendre condition for \(\lambda = \frac{1}{4}\). Thus, we can solve (3.8) and (3.9) with homogeneous Dirichlet boundary condition

$$\begin{aligned} \begin{pmatrix} b_2 \\ c_2 \end{pmatrix} = 0 \quad \text {and}\quad \begin{pmatrix} b_3 \\ c_3 \end{pmatrix} = 0 \quad \text {on}\quad \partial S \end{aligned}$$

as the system (3.9) can be treated in the same manner. The regularity of a implies now that \(b = (b_2,b_3)\) and \(c=(c_2,c_3)\) are \(C^\infty ({\overline{S}};\mathbb {R}^2)\).

The approximating solution \(\tilde{u}_h\) solves then the following system

$$\begin{aligned}&\partial _t^2 \tilde{u}_h - \frac{1}{h^2} {\text {div}}_h\Big (D^2\tilde{W}(0) \nabla _h\tilde{u}_h\Big ) = h^{2} f_h - r_h \quad \text {in}\quad \Omega \times (0,T),\\&D^2\tilde{W}(0)[\nabla _h \tilde{u}_h]\nu \Big |_{(0,L)\times \partial S} = {\text {tr}}_{\partial \Omega }(r_{N,h}) \nu \quad \text {on}\quad \partial \Omega \times (0,T),\\&\tilde{u}_h\,\text {is}\,L\text {-periodic}\quad \text {in}\quad x_1\text {-direction},\\&(\tilde{u}_h, \partial _t \tilde{u}_h)|_{t=0} = (\tilde{u}_{0,h}, \tilde{u}_{1,h}), \end{aligned}$$

where \(r_h\) is chosen as above,

$$\begin{aligned} r_{N,h} := h^5 \begin{pmatrix} 0 \\ \nu ^T \begin{pmatrix} \partial _{x_2} b_2 &{} \frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2) \\ \partial _{x_2} c_3 &{} \frac{1}{2} (\partial _{x_2} b_3 + \partial _{x_3} c_3) \end{pmatrix} \partial _{x_1}^4 v \\ \nu ^T \begin{pmatrix} \frac{1}{2}(\partial _{x_2} c_2 + \partial _{x_3} b_2) &{} \partial _{x_3} c_2 \\ \frac{1}{2}(\partial _{x_2} b_3 + \partial _{x_3} c_3) &{} \partial _{x_3} b_3 \end{pmatrix}\partial _{x_1}^4 v \end{pmatrix}, \end{aligned}$$

and the initial data is given by

$$\begin{aligned} \tilde{u}_{j,h}(x,t)&= h^2 \begin{pmatrix} 0 \\ v^j_2 \\ v^j_3 \end{pmatrix} + h^3 \begin{pmatrix} -x_2 \partial _{x_1} v^j_{2} - x_3 \partial _{x_1} v^j_{3} \\ 0 \\ 0 \end{pmatrix} + h^5 \begin{pmatrix} a_2(x') \partial _{x_1}^3 v^j_2 + a_3(x') \partial _{x_1}^3 v^j_3 \\ 0 \\ 0 \end{pmatrix}\nonumber \\&\quad + h^6 \begin{pmatrix} 0 \\ b_2(x') \partial _{x_1}^4 v^j_2 + c_3(x') \partial _{x_1}^4 v^j_3 \\ b_3(x') \partial _{x_1}^4 v^j_3 + c_2(x') \partial _{x_1}^4 v^j_2 \end{pmatrix} \end{aligned}$$
(3.10)

with \(v^j := \partial _t^j v|_{t=0}\) and \(j=0,\ldots , 4\). For the remainder it holds

$$\begin{aligned} \Vert r_h\Vert _{C^0([0,T];L^2)} \le Ch^3\quad \text {and}\quad \Vert r_{N,h}\Vert _{C^2([0,T]; H^1)} \le Ch^5. \end{aligned}$$

3.2 Existence of and bounds on initial values

Define now

$$\begin{aligned} \mathcal {B} := H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\cap \Big \{u\in L^2(\Omega ;\mathbb {R}^3) \;:\; \int \limits _\Omega udx = \int \limits _\Omega u\cdot x^\perp dx = 0 \Big \} \end{aligned}$$

equipped with the norm

$$\begin{aligned} \Vert u\Vert _{\mathcal {B}_h} := \bigg \Vert \frac{1}{h} \varepsilon _h(u)\bigg \Vert _{L^2(\Omega )}. \end{aligned}$$

Lemma 3.1

There exist constants \(C_0>0\) and \(M_0\in (0,1]\) such that for \(0<h\le 1\) and \(f\in H^{1,1}_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) with \(\Vert f\Vert _{H^{1,1}(\Omega )} \le M_0 h\) and \(\int \limits _\Omega f dx = 0\) there exists a unique solution \(w\in H^3_{\textrm{per}}(\Omega ; \mathbb {R}^3)\cap \mathcal {B}\) with \(\partial _{x_1} w\in H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) of

$$\begin{aligned} \frac{1}{h^2}\Big (D\tilde{W}(\nabla _h w), \nabla _h \varphi \Big )_{L^2(\Omega )} = (f,\varphi )_{L^2(\Omega )}\quad \text {for all}\quad \varphi \in \mathcal {B}. \end{aligned}$$
(3.11)

Moreover

$$\begin{aligned} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(w), \nabla _h\frac{1}{h}\varepsilon (w), \nabla _h^2 w\bigg )\bigg \Vert _{H^{1,1}(\Omega )} \le C_0 \Vert f\Vert _{H^{1,1}(\Omega )} \end{aligned}$$
(3.12)

holds. If \(w'\in H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\cap \mathcal {B}\) with \(\partial _{x_1} w\in H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) is the solution to \(f'\in H^{1,1}_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) with \(\Vert f'\Vert _{H^{1,1}(\Omega )} \le M_0 h\) and \(\int \limits _\Omega f' dx = 0\), then it holds

$$\begin{aligned} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(w - w'), \nabla _h\frac{1}{h}\varepsilon (w - w'), \nabla _h^2 (w - w')\bigg )\bigg \Vert _{H^{1,1}(\Omega )} \le C_0 \Vert f - f'\Vert _{H^{1,1}(\Omega )}. \end{aligned}$$
(3.13)

Proof

Using a Taylor series expansion for \(D\tilde{W}(\nabla _h w)\), we obtain

$$\begin{aligned} D\tilde{W}(\nabla _h w)&= D\tilde{W}(0) + D^2\tilde{W}(0)[\nabla _h w] + \int \limits _0^1 (1-\tau ) D^3\tilde{W}(\tau \nabla _h w)[\nabla _h w, \nabla _h w]d\tau \nonumber \\&=: D^2\tilde{W}(0)\nabla _h w + G(\nabla _h w). \end{aligned}$$
(3.14)

Thus, (3.11) is equivalent to

$$\begin{aligned} \langle L_h w, \varphi \rangle _{\mathcal {B}', \mathcal {B}} :=\frac{1}{h^2}\Big (D^2\tilde{W}(0)\nabla _h w, \nabla _h \varphi \Big )_{L^2(\Omega )} = (f,\varphi )_{L^2(\Omega )} - \frac{1}{h^2}(G(\nabla _h w), \nabla _h \varphi )_{L^2(\Omega )}. \end{aligned}$$

The idea is now to use the contraction mapping principle in order to prove the existence of a solution for (3.11), i.e., with the later equivalence

$$\begin{aligned} w = \mathcal {G}_{h,f} (w) := L_h^{-1}\Big (f, G_h(w)\Big ) \end{aligned}$$

holds with \(G_h(w) := \frac{1}{h^2} G(\nabla _h w)\). Consequently, we investigate the mapping properties of \(L_h\) and \(G_h\).

For \(f\in L^2(\Omega ;\mathbb {R}^3)\) and \(F\in L^2(\Omega ;\mathbb {R}^{3\times 3})\), we obtain with the Lemma of Lax–Milgram the existence of a unique solution \(w\in \mathcal {B}\) for

$$\begin{aligned} \langle L_h w, \varphi \rangle _{\mathcal {B}',\mathcal {B}} = (f,\varphi )_{L^2(\Omega )} - (F, \nabla _h \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.15)

for all \(\varphi \in \mathcal {B}\). The solution satisfies

$$\begin{aligned} \Vert w\Vert _{\mathcal {B}} = \bigg \Vert \frac{1}{h} \varepsilon _h(w)\bigg \Vert _{L^2(\Omega )} \le C \big ( \Vert f\Vert _{L^2(\Omega )} + \Vert F\Vert _{(L^2_h)'}\big ). \end{aligned}$$

If now \(f\in H^{0,k}(\Omega ;\mathbb {R}^3)\) and \(F\in H^{0,k}(\Omega ;\mathbb {R}^{3\times 3})\) for \(k=1,2\), it follows by a different quotient argument that \(w\in H^{0,k}(\Omega ;\mathbb {R}^3)\) holds and

$$\begin{aligned} \bigg \Vert \frac{1}{h}\varepsilon _h(w)\bigg \Vert _{H^{0,k}(\Omega )} \le C \Big (\Vert f\Vert _{H^{0,k-1}(\Omega )} + \max _{j=0,\ldots , k}\Vert \partial _{x_1}^j F\Vert _{(L^2_h)'}\Big ). \end{aligned}$$
(3.16)

Using the decomposition \(\mathcal {B}\oplus {\text {span}}\{x\mapsto x^\perp \} = H^1_{(0),per}(\Omega ;\mathbb {R}^3)\), it follows that for

$$\begin{aligned} \alpha := (F, \nabla _h x^\perp )_{L^2(\Omega )} - (f,x^\perp )_{L^2(\Omega )} \end{aligned}$$

we have

$$\begin{aligned} \frac{1}{h^2}\Big (D^2\tilde{W}(0)\nabla _h w, \nabla _h \varphi \Big )_{L^2(\Omega )} = (f + \alpha x^\perp ,\varphi )_{L^2(\Omega )} - (F, \nabla _h \varphi )_{L^2(\Omega )} \end{aligned}$$

for all \(\varphi \in H^1_{(0),\textrm{per}}(\Omega ;\mathbb {R}^3)\). Hence, if \(f\in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) and \(F\in H^2_{\textrm{per}}(\Omega ;\mathbb {R}^{3\times 3})\), then w solves the system

$$\begin{aligned} \left\{ \begin{array}{ll} - \frac{1}{h^2} {\text {div}}_h (D^2\tilde{W}(0)\nabla _h w) = f + \alpha x^\perp - {\text {div}}_h F &{}\quad \text {in}\quad \Omega \\ D^2\tilde{W}(0)[\nabla _h w]\nu \Big |_{\partial S} = h^2 {\text {tr}}_{\partial \Omega }(F)\nu \Big |_{\partial S} &{}\quad \text {in}\quad \partial \Omega \end{array}\right. \end{aligned}$$

in a weak sense. Thus, with elliptic regularity theory it follows \(w\in H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\cap \mathcal {B}\). By Theorem A.3 in the appendix, we obtain

$$\begin{aligned} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(w), \nabla \frac{1}{h}\varepsilon _h(w), \nabla _h^2 w\bigg )\bigg \Vert _{H^1(\Omega )}&\le C\bigg (h^2\Vert (f,{\text {div}}_h F)\Vert _{H^1(\Omega )} + \Vert f\Vert _{H^{0,1}(\Omega )}\\&\quad + \max _{j=0,1,2}\Vert \partial _{x_1}^j F\Vert _{(L^2_h)'} \\&\quad + \Big \Vert h {\text {tr}}_{\partial \Omega }(F) \Big \Vert _{L^2(0,L; H^\frac{3}{2}(\partial S))\cap H^1(0,L;H^\frac{1}{2}(\partial S))} \bigg ), \end{aligned}$$

where we have exploited

$$\begin{aligned} h^2 |\alpha | \le Ch^2 \Vert f\Vert _{L^2(\Omega )} + Ch \Vert F\Vert _{(L^2_h)'}. \end{aligned}$$

Using that \({\text {tr}}_{\partial S}:H^2(S) \rightarrow H^{\frac{3}{2}}(\partial S)\) is a bounded operator, we obtain

$$\begin{aligned} h \Big \Vert {\text {tr}}_{\partial \Omega }(F) \Big \Vert _{L^2(0,L; H^\frac{3}{2}(\partial S))\cap H^1(0,L;H^\frac{1}{2}(\partial S))}&\le Ch \Big (\Vert F\Vert _{H^{1,1}(\Omega )} + \max _{k=0,1,2}\Vert \nabla _{x'}^k F\Vert _{L^2(\Omega )} \Big )\\&\le C\Big (\max _{j=0,1,2}\Vert \partial _{x_1}^j F\Vert _{(L^2_h)'} + h^2 \Vert \nabla _h F\Vert _{H^1(\Omega )}\Big ) \end{aligned}$$

because of

$$\begin{aligned} \Vert F\Vert _{H^1(\Omega )} \le \Vert F\Vert _{H^{0,1}(\Omega )} + \Vert \nabla _{x'} F\Vert _{L^2(\Omega )}\;\text { and }\; \Vert F\Vert _{L^2(\Omega )} \le \frac{1}{h} \Vert F\Vert _{(L^2_h)'}. \end{aligned}$$

Thus, we deduce for some \(C_L > 0\)

$$\begin{aligned}&\bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(w), \nabla \frac{1}{h}\varepsilon _h(w), \nabla _h^2 w\bigg )\bigg \Vert _{H^1(\Omega )}\nonumber \\&\quad \le C_L \Big (h^2\Vert (f, \nabla _h F)\Vert _{H^1(\Omega )} + \Vert f\Vert _{H^{0,1}(\Omega )} + \max _{j=0,1,2}\Vert \partial _{x_1}^j F\Vert _{(L^2_h)'}\Big ). \end{aligned}$$
(3.17)

We define \(\mathcal {X}_h := H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\cap \mathcal {B}\) and \(\mathcal {Y}_h := H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3) \times H^2_{\textrm{per}}(\Omega ;\mathbb {R}^{3\times 3})\) normed via

$$\begin{aligned} \Vert g\Vert _{\mathcal {X}_h}:= & {} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(g), \nabla \frac{1}{h}\varepsilon _h(g), \nabla _h^2 g\bigg )\bigg \Vert _{H^1(\Omega )}\\ \Vert (f,F)\Vert _{\mathcal {Y}_h}:= & {} h^2\Vert (f, \nabla _h F)\Vert _{H^1(\Omega )} + \Vert f\Vert _{H^{0,1}(\Omega )} + \max _{j=0,1,2}\Vert \partial _{x_1}^j F\Vert _{(L^2_h)'}. \end{aligned}$$

This \(L_h^{-1}:\mathcal {Y}_h \rightarrow \mathcal {X}_h\) is a bilinear, bijective and bounded operator, mapping a tuple \((f,F)\in \mathcal {Y}_h\) to the corresponding solution \(w\in \mathcal {X}_h\) of (3.15). In order to close the proof, we have to show that \(G_h\) is a contraction with respect to the relevant norms.

In a first step we assume that \(w_i\in \mathcal {X}_h\) with

$$\begin{aligned} \Vert w_i\Vert _{\mathcal {X}_h} \le C_0 M_1 h \end{aligned}$$

for \(i=1,2\) and \(M_1 > 0\) to be chosen later. Then

$$\begin{aligned} \Vert G_h (w_1) - G_h(w_2)\Vert _{(L^2_h)'}&= \bigg \Vert \frac{1}{h^2} \int \limits _0^1 (1-\tau ) \Big (D^3\tilde{W}(\tau \nabla _h w_1)[\nabla _h w_1 - \nabla _h w_2, \nabla _h w_1]\\&\qquad + D^3\tilde{W}(\tau \nabla _h w_2)[\nabla _h w_1 - \nabla _h w_2, \nabla _h w_2]\\&\qquad + \big (D^3\tilde{W}(\tau \nabla _h w_1) - D^3\tilde{W}(\tau \nabla _h w_2)\big )[\nabla _h w_1, \nabla _h w_2]\Big ) d\tau \bigg \Vert _{(L^2_h)'}\\&\quad \le CM_1 \Vert \nabla _h (w_1 - w_2)\Vert _{H^1_h(\Omega )}\\&\qquad + \bigg \Vert \frac{1}{h^2} \int \limits _0^1(1-\tau ) \int \limits _0^1 Q(\tau , t, w_1, w_2) dt [\tau (\nabla _h w_1 - \nabla _h w_2), \nabla _h w_1, \nabla _h w_2]d\tau \bigg \Vert _{(L^2_h)'}\\&\quad \le C M_1 \bigg \Vert \frac{1}{h}\varepsilon _h(w_1 - w_2)\bigg \Vert _{\mathcal {X}_h}, \end{aligned}$$

where we used Corollary 2.4, \(\Vert \nabla _h w_j\Vert _{H^1_h(\Omega )}\le C\Vert w_j\Vert _{\mathcal {X}_h}\) and the boundedness of

$$\begin{aligned} Q(\tau , t, w_1, w_2) := D^4\tilde{W}(t\tau \nabla _h w_1 + (1-t)\tau \nabla _h w_2). \end{aligned}$$

The definition of G implies that for \(k=1,2,3\) it holds

$$\begin{aligned} \partial _{x_k} G(\nabla _h w)&= D^2\tilde{W}(\nabla _h w)[\nabla _h \partial _{x_k} w] - D^2\tilde{W}(0) [\nabla _h \partial _{x_k} w] \nonumber \\&= \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h w) [\nabla _h w, \nabla _h\partial _{x_k} w] d\tau . \end{aligned}$$
(3.18)

Hence, analogously as above

$$\begin{aligned}&\Vert \partial _{x_k} (G_h(w_1) - G_h(w_2)\Vert _{(L^2_h)'} \le \bigg \Vert \frac{1}{h^2} \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h w_1) [\nabla _h (w_1 - w_2), \nabla _h \partial _{x_k} w_1]d\tau \bigg \Vert _{(L^2_h)'}\\&\qquad + \bigg \Vert \frac{1}{h^2} \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h w_2) [\nabla _h w_2, \nabla _h \partial _{x_k} (w_1 - w_2)]d\tau \bigg \Vert _{(L^2_h)'}\\&\qquad + \bigg \Vert \frac{1}{h^2} \int \limits _0^1 \big (D^3\tilde{W}(\tau \nabla _h w_1) - D^3\tilde{W}(\tau \nabla _h w_2)\big ) [\nabla _h w_2, \nabla _h \partial _{x_k} w_1]d\tau \bigg \Vert _{(L^2_h)'}\\&\quad \le \frac{C}{h}\Vert \nabla _h (w_1 - w_2)\Vert _{H^2_h(\Omega )} \Vert \nabla _h \partial _{x_k} w_1\Vert _{L^2_h(\Omega )} + \frac{C}{h} \Vert \nabla _h w_2\Vert _{H^2_h(\Omega )} \Vert \nabla _h \partial _{x_k}(w_1 - w_2)\Vert _{(L^2_h)'}\\&\qquad + CM_1 \bigg \Vert \frac{1}{h}\varepsilon _h(w_1 - w_2)\bigg \Vert _{H^1_h(\Omega )}\\&\quad \le CM_1 \Vert w_1 - w_2\Vert _{\mathcal {X}_h} \end{aligned}$$

as

$$\begin{aligned} \Vert \nabla _h \partial _{x_k} \varphi \Vert _{L^2_h(\Omega )} \le \Vert \nabla _h \partial _{x_k} \varphi \Vert _{L^2(\Omega )} + \bigg \Vert \frac{1}{h} \varepsilon _h(\partial _{x_k} \varphi )\bigg \Vert _{L^2(\Omega )} \le \bigg \Vert \bigg (\nabla \frac{1}{h}\varepsilon _h(\varphi ), \nabla _h^2 \varphi \bigg )\bigg \Vert _{L^2(\Omega )} \le \Vert \varphi \Vert _{\mathcal {X}_h} \end{aligned}$$

for \(\varphi = w_1\) and \(\varphi = w_1 - w_2\). With the aid of (3.18) it follows for \(j, k=1, 2, 3\)

$$\begin{aligned} \partial _{x_j}\partial _{x_k} G(\nabla _h w)&= D^2\tilde{W}(\nabla _h u_h)[\nabla _h\partial _{x_j}\partial _{x_k} w] - D^2\tilde{W}(0)[\nabla _h\partial _{x_j}\partial _{x_k} w] \\&\quad + D^3\tilde{W}(\nabla _h w) [\nabla _h \partial _{x_j} w, \nabla _h \partial _{x_k} w]\\&= \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h w)[\nabla _h w, \nabla _h\partial _{x_j}\partial _{x_k} w] d\tau + D^3\tilde{W}(\nabla _h w) [\nabla _h \partial _{x_j} w, \nabla _h \partial _{x_k} w]. \end{aligned}$$

Thus, we obtain in the same manner as above

$$\begin{aligned} \Vert \partial _{x_j}\partial _{x_k} (G_h(w_1) - G_h(w_2))\Vert _{(L^2_h)'} \le CM_1 \Vert w_1 - w_2\Vert _{\mathcal {X}_h}. \end{aligned}$$

The fact that \(h^2\Vert \nabla _h F\Vert _{H^1(\Omega )} \le h \Vert \nabla F\Vert _{H^1(\Omega )}\) and \(\Vert F\Vert _{L^2(\Omega )} \le \frac{1}{h}\Vert F\Vert _{(L^2_h)'}\) implies with the later estimates that for \(M_1\in (0,1]\) small enough

$$\begin{aligned} \mathcal {G}_{h,f}:\overline{B_{CM_1 h}(0)} \subset \mathcal {X}_h \rightarrow \mathcal {X}_h \end{aligned}$$

is a \(\frac{1}{2}\)-contraction. The self-mapping property of \(\mathcal {G}_{h,f}\) follows because of

$$\begin{aligned} \Vert \mathcal {G}_{h,f} (0)\Vert _{\mathcal {X}_h} = \Vert L^{-1}(f,0)\Vert _{\mathcal {X}_h} \le C_L \Vert (f,0)\Vert _{\mathcal {Y}_h} \le C_L \Vert f\Vert _{H^{1,1}(\Omega )} \le C_L M_0 h. \end{aligned}$$

Thus, we can choose \(M_0 > 0\) so small that \(C_L M_0 h \le \frac{C M_1}{2}\). Then we obtain with the \(\frac{1}{2}\)-contraction property of \(\mathcal {G}_{h,f}\) for \(w\in \overline{B_{CM_1 h}(0)}\)

$$\begin{aligned} \Vert \mathcal {G}_{h,f}(w)\Vert _{\mathcal {X}_h} \le \Vert \mathcal {G}_{h,f}(w) - \mathcal {G}_{h,f}(0)\Vert _{\mathcal {X}_h} + \Vert \mathcal {G}_{h,f}(0)\Vert _{\mathcal {X}_h} \le \frac{1}{2} \Vert w\Vert _{\mathcal {X}_h} + C_L M_0 h \le C M_1 h. \end{aligned}$$

Therefore, (3.12) and (3.13) hold with the \(H^{1,1}(\Omega )\)-norm on the left-hand side replaced by the \(\mathcal {X}_h\)-norm.

Using the decomposition \(\mathcal {B}\oplus {\text {span}}\{x\mapsto x^\perp \} = H^1_{(0),per}(\Omega ;\mathbb {R}^3)\), it follows that for

$$\begin{aligned} \rho := \frac{1}{\mu (S) h^2} \Big (D\tilde{W}(\nabla _h w), \nabla _h x^\perp \Big )_{L^2(\Omega )} \end{aligned}$$

we have

$$\begin{aligned} \frac{1}{h^2}\Big (D\tilde{W}(\nabla _h w), \nabla _h \varphi \Big )_{L^2(\Omega )} = (f - \rho x^\perp ,\varphi )_{L^2(\Omega )} \end{aligned}$$

for all \(\varphi \in H^1_{(0),per}(\Omega ;\mathbb {R}^3)\). If now \(f\in H^{1,1}_{\textrm{per}}(\Omega ;\mathbb {R}^3)\) we obtain, with a difference quotient argument, that \(w\in H^3_{\textrm{per}}(\Omega ;\mathbb {R}^3)\cap \mathcal {B}\) satisfies

$$\begin{aligned} \frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h w) \nabla _h\partial _{x_1} w, \nabla _h \varphi \Big )_{L^2(\Omega )} = (\partial _{x_1} f, \varphi )_{L^2(\Omega )}. \end{aligned}$$

for all \(\varphi \in H^1_{(0),per}(\Omega ;\mathbb {R}^3)\). Thus, with Theorem A.3 the claimed inequalities follow. \(\square \)

We define the initial values for the analytical problem as

$$\begin{aligned} u_{2+j, h} :=h^{2} \begin{pmatrix} 0 \\ v^{2+j}_2 \\ v^{2+j}_3 \end{pmatrix} + h^{3} \begin{pmatrix} -x_2 \partial _{x_1}v^{2+j}_{2} - x_3 \partial _{x_1}v^{2+j}_{3} \\ 0 \\ 0 \end{pmatrix} \end{aligned}$$

for \(j=1,2\) and \(v^{2+j} = \partial _t^{2+j} v|_{t=0}\) as above.

Lemma 3.2

Let \(\tilde{u}_h\) be as in (3.7), \(\tilde{u}_{j,h}\) for \(j=0,1,2\) as in (3.10), \(u_{3,h}\), \(u_{4,h}\) and \(f^h\) be as above. Then for sufficiently small \(h_0\in (0,1]\) and \(M>0\), there exist solutions \((u_{0,h}, u_{1,h}, u_{2,h})\) of

$$\begin{aligned} \frac{1}{h^2} \Big (D\tilde{W}(\nabla _h u_{0,h}), \nabla _h\varphi \Big )_{L^2(\Omega )}&= (h^2 f_h|_{t=0}, \varphi )_{L^2(\Omega )} - (u_{2,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.19)
$$\begin{aligned} \frac{1}{h^2} \Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{1,h}, \nabla _h\varphi \Big )_{L^2(\Omega )}&= (h^2 \partial _t f|_{t=0}, \varphi )_{L^2(\Omega )} - (u_{3,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.20)

and

$$\begin{aligned}&\frac{1}{h^2} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) \nabla _h u_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2\partial _t^2 f|_{t=0} - u_{4,h})_{L^2(\Omega )}\nonumber \\&\quad - \frac{1}{h^2}\Big (D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h\varphi \Big )_{L^2(\Omega )} - \frac{\gamma _h}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h})P,\nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$
(3.21)

for all \(\varphi \in \mathcal {B}\), where

$$\begin{aligned} \gamma _h(u_{0,h}) := \frac{1}{\mu (S)h^3} \Big (D\tilde{W}(\nabla _h u_{0,h}), P\Big )_{L^2(\Omega )} \end{aligned}$$

and

$$\begin{aligned} P:= \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad -1\\ 0 &{}\quad 1 &{}\quad 0 \end{pmatrix}. \end{aligned}$$

The solution satisfies

$$\begin{aligned} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h (u_{0,h}), \nabla \frac{1}{h}\varepsilon _h (u_{0,h}), \nabla _h^2 u_{0,h}\bigg )\bigg \Vert _{H^{1,1}(\Omega )} \le Ch^2 \end{aligned}$$
(3.22)
$$\begin{aligned} \max _{j=1,2} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h (u_{j,h}), \nabla \frac{1}{h}\varepsilon _h (u_{j,h}), \nabla _h^2 u_{j,h}\bigg )\bigg \Vert _{H^{2-j}(\Omega )}\le Ch^2 \end{aligned}$$
(3.23)

and \(u_{k,h}\in \mathcal {B}\) for \(k=0,1,2\). Moreover, we have

$$\begin{aligned} \max _{j=0,1,2} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(u_{j,h}) - \frac{1}{h}\varepsilon _h(\tilde{u}_{j,h})\bigg )\bigg \Vert _{L^2(\Omega )} \le {\left\{ \begin{array}{ll} Ch^3 &{}\text {if}\quad j=0,1,\\ Ch^2 &{}\text {if}\quad j=2 \end{array}\right. } \end{aligned}$$
(3.24)

for all \(h\in (0,h_0]\) and \(C>0\) independent of h.

Proof

We can equivalently formulate (3.19)–(3.21) via

$$\begin{aligned} \frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{1,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2\partial _t f_h|_{t=0}, \varphi )_{L^2(\Omega )} - (u_{3,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.25)

and

$$\begin{aligned}&\frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2\partial _t^2 f_h|_{t=0} - u_{4,h}, \varphi )_{L^2(\Omega )}\nonumber \\&\quad -\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h}) [\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} \nonumber \\&\quad - \frac{\gamma _h(u_{0,h})}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) P,\nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$
(3.26)

for all \(\varphi \in \mathcal {B}\), where \(u_{0,h} = \mathcal {G}_{h,f}(u_{0,h})\) is the solution of (3.19) with \(f = h^2 f^h - u_{2,h}\). Defining

$$\begin{aligned} \mathcal {G}_{0,h}(u_{2,h}) := \mathcal {G}_{h,f}(u_{0,h}) \end{aligned}$$

and deploying (3.13) we obtain for \(u_{2,h}\), \(u'_{2,h} \in H^{1,1}(\Omega ;\mathbb {R}^3)\)

$$\begin{aligned} \max _{k=0,1}\Big \Vert \partial _{x_1}^k \big (\mathcal {G}_{0,h}(u_{2,h}) - \mathcal {G}_{0,h}(u'_{2,h})\big )\Big \Vert _{\mathcal {X}_h} \le C_0 \Vert u_{2,h} - u'_{2,h}\Vert _{H^{1,1}(\Omega )} \end{aligned}$$
(3.27)

if \(\Vert u_{2,h}\Vert _{H^{1,1}(\Omega )} \le \frac{1}{2}M_0 h\), \(\Vert u'_{2,h}\Vert _{H^{1,1}(\Omega )} \le \frac{1}{2}M_0 h\) and \(h^2\Vert f^h\Vert _{H^{1,1}(\Omega )}\le \frac{1}{2}M_0 h\). This can always be achieved if \(h_0\in (0,1]\) is small enough and \(u_{2,h}\), \(u'_{2,h}\) are of order \(h^2\).

Using the definition of \(L_h\) it follows that (3.25)–(3.26) are equivalent to

$$\begin{aligned} \langle L_h u_{1,h}, \varphi \rangle _{\mathcal {B}', \mathcal {B}} = (h^2\partial _t f_h|_{t=0} - u_{3,h}, \varphi )_{L^2(\Omega )} - \frac{1}{h^2} \Big (DG(\nabla _h u_{0,h}) \nabla _h u_{1,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$

and

$$\begin{aligned} \langle L_h u_{2,h}, \varphi \rangle _{\mathcal {B}', \mathcal {B}}&= (h^2\partial ^2_t f_h|_{t=0} - u_{4,h}, \varphi )_{L^2(\Omega )} - \frac{1}{h^2} \Big (DG(\nabla _h u_{0,h}) \nabla _h u_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )}\\&\quad - \frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} \\&\quad - \frac{\gamma _h(u_{0,h})}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h})P,\nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$

for all \(\varphi \in \mathcal {B}\). Defining now the relevant function spaces by

$$\begin{aligned} \mathcal {D}_h&:= H^2_{\textrm{per}}(\Omega ;\mathbb {R}^{3\times 3})\times H^1_{\textrm{per}}(\Omega ;\mathbb {R}^{3\times 3}), \quad \mathcal {Z}_h := H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\times L^2 (\Omega ;\mathbb {R}^3)\times \mathcal {D}_h,\\ \mathcal {W}_h&:= \mathcal {X}_h \times \Big (H^2_{\textrm{per}}(\Omega ;\mathbb {R}^3) \cap \mathcal {B}\Big ) \end{aligned}$$

with the respective norms defined by

$$\begin{aligned} \Vert (F_1, F_2)\Vert _{\mathcal {D}_h}&:= \max _{i=1,2} \Big (h^2\Vert \nabla _h F_i\Vert _{H^{2-i}(\Omega )} + \max _{\sigma =0,\ldots , 3-i}\Vert \partial _{x_1}^\sigma F_i\Vert _{(L^2_h)'}\Big ),\\ \Vert (f_1,f_2, F_1, F_2)\Vert _{\mathcal {Z}_h}&:= \max _{i=1,2} \Big (h^2\Vert (f_i, \nabla _h F_i)\Vert _{H^{2-i}(\Omega )} + \Vert f_i\Vert _{H^{0,2-i}(\Omega )} + \max _{\sigma =0,\ldots , 3-i}\Vert \partial _{x_1}^\sigma F_i\Vert _{(L^2_h)'}\Big ),\\ \Vert (g_1, g_2)\Vert _{\mathcal {W}_h}&:= \max _{i=1,2} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(g_i), \nabla \frac{1}{h}\varepsilon _h(g_i),\nabla _h^2 g_i\bigg )\bigg \Vert _{H^{2-i}(\Omega )}. \end{aligned}$$

With this we define the linear operator \(\mathcal {L}_h^{-1}:\mathcal {Z}_h \rightarrow \mathcal {W}_h\) by mapping \((f_1,f_2, F_1, F_2)\) to the solution \((w_1, w_2)\) of

$$\begin{aligned} \langle L_h w_i, \varphi \rangle _{\mathcal {B}',\mathcal {B}} = (f_i,\varphi )_{L^2(\Omega )} - (F_i, \nabla _h \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.28)

for \(i=1,2\). Then due to (3.17), Theorem A.3 and (3.16) we obtain

$$\begin{aligned} \Vert (w_1, w_2)\Vert _{\mathcal {W}_h}\le C \Vert (f_1,f_2,F_1,F_2)\Vert _{\mathcal {Z}_h}. \end{aligned}$$
(3.29)

Hence, \(\mathcal {L}_h^{-1}\) is a bijective, linear and bounded operator. For the nonlinearity we define

$$\begin{aligned} \mathcal {Q}_h:\mathcal {W}_h\rightarrow \mathcal {D}_h \end{aligned}$$

via

$$\begin{aligned} \begin{pmatrix} u_1 \\ u_2 \end{pmatrix}&\mapsto \begin{pmatrix} -\frac{1}{h^2} DG(\nabla _h u_0) \nabla _h u_1\\ -\frac{1}{h^2} DG(\nabla _h u_0) \nabla _h u_2 - \frac{1}{h^2} D^3\tilde{W}(\nabla _h u_0) [\nabla _h u_1, \nabla _h u_1] - \frac{\gamma _h(u_{0,h})}{h^3} D^2\tilde{W}(\nabla _h u_{0,h})[P] \end{pmatrix} \\&=: \begin{pmatrix} \mathcal {Q}_{1,h}(u_1, u_2)\\ \mathcal {Q}_{2,h}(u_1, u_2) \end{pmatrix}, \end{aligned}$$

where \(u_0 := \mathcal {G}_{h,f - u_2}(u_{0,h})\) for some fixed \(f\in H^{1,1}_{\textrm{per}}(\Omega )\) with \(\Vert f\Vert _{H^{1,1}(\Omega )} \le M h^2\) and \(\int \limits _\Omega f dx = 0\) and G is defined as in (3.14).

We deduce the contraction properties of \(\mathcal {Q}_h\) similar as in the proof of Lemma 3.1. For this we assume that \(\Vert (u_1, u_2)\Vert _{\mathcal {W}_h}\), \(\Vert (u'_1, u'_2)\Vert _{\mathcal {W}_h}\le CM_2 h^2\). Starting with \(\mathcal {Q}_{1,h}\), we obtain

$$\begin{aligned}&\Vert \mathcal {Q}_{1,h}(u_1, u_2) - \mathcal {Q}_{1,h}(u'_1, u'_2)\Vert _{(L^2_h)'}\\&\quad = \frac{1}{h^2} \bigg \Vert \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u_0)[\nabla _h u_0, \nabla _h u_1] d\tau - \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u'_0)[\nabla _h u'_0, \nabla _h u'_1]d\tau \bigg \Vert _{(L^2_h)'}\\&\quad \le \frac{C}{h}\Vert \nabla _h (u_0 - u'_0)\Vert _{H^2_h(\Omega )} \Vert \nabla _h u_1\Vert _{L^2_h(\Omega )} + \frac{C}{h} \Vert \nabla _h u'_0\Vert _{H^2_h(\Omega )} \Vert \nabla _h (u_1 - u'_1)\Vert _{L^2_h(\Omega )}\\&\qquad + CM_2 \bigg \Vert \frac{1}{h}\varepsilon _h(u_0 - u'_0)\bigg \Vert _{H^1_h(\Omega )}\\&\quad \le CM_2 \Vert u_2 - u'_2\Vert _{H^{1,1}(\Omega )} + CM_2 \bigg \Vert \frac{1}{h}\varepsilon _h(u_1 - u'_1)\bigg \Vert _{L^2(\Omega )} \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h}, \end{aligned}$$

where we used (3.27). Similarly one deduces that

$$\begin{aligned} \Vert \partial _{x_j} (\mathcal {Q}_{1,h}(u_1, u_2) - \mathcal {Q}_{1,h}(u'_1, u'_2))\Vert _{L^2(\Omega )} \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h}\\ \Vert \partial _{x_k} \partial _{x_j} (\mathcal {Q}_{1,h}(u_1, u_2) - \mathcal {Q}_{1,h}(u'_1, u'_2))\Vert _{L^2(\Omega )} \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h} \end{aligned}$$

for \(j,k=1,2,3\). Analogously we deduce for \(\mathcal {Q}_{2,h}\)

$$\begin{aligned}&\Vert \mathcal {Q}_{2,h}(u_1, u_2) - \mathcal {Q}_{2,h}(u'_1, u'_2)\Vert _{(L^2_h)'}\\&\quad \le \frac{1}{h^2}\bigg \Vert \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u_0)[\nabla _h u_0, \nabla _h u_2] - D^3\tilde{W}(\tau \nabla _h u'_0)[\nabla _h u'_0, \nabla _h u'_2]d\tau \bigg \Vert _{(L^2_h)'}\\&\qquad + \frac{|\gamma _h(u_{0,h})|}{h^3}\bigg \Vert \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u_0)[\nabla _h u_0, P] - D^3\tilde{W}(\tau \nabla _h u'_0)[\nabla _h u'_0, P]d\tau \bigg \Vert _{(L^2_h)'}\\&\qquad + \frac{|\gamma _h(u_{0,h}) - \gamma _h (u_{0,h}')|}{h^3} \bigg \Vert \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u'_0)[\nabla _h u'_0, P]d\tau \bigg \Vert _{(L^2_h)'}\\&\qquad + \frac{1}{h^2} \bigg \Vert D^3\tilde{W}(\nabla _h u_0)[\nabla _h u_1, \nabla _h u_1] - D^3\tilde{W}(\nabla _h u'_0)[\nabla _h u'_1, \nabla _h u'_1]\bigg \Vert _{(L^2_h)'}\\&\quad \le \frac{C}{h}\Vert \nabla _h (u_0 - u'_0)\Vert _{H^2_h(\Omega )} \Vert \nabla _h u_2\Vert _{L^2_h(\Omega )} + \frac{C}{h} \Vert \nabla _h u'_0\Vert _{H^2_h(\Omega )} \Vert \nabla _h (u_2 - u'_2)\Vert _{L^2_h(\Omega )}\\&\qquad + \frac{C}{h}\Vert \nabla _h (u_1 - u'_1)\Vert _{H^2_h(\Omega )} \Vert \nabla _h u_1\Vert _{L^2_h(\Omega )} + \frac{C}{h} \Vert \nabla _h u'_1\Vert _{H^2_h(\Omega )} \Vert \nabla _h (u_1 - u'_1)\Vert _{L^2_h(\Omega )}\\&\qquad + \frac{C}{h^2} \Vert \nabla _h(u_0 - u'_0)\Vert _{H^2_h(\Omega )} \Vert \nabla _h u_1\Vert _{H^1_h(\Omega )} \Vert \nabla _h u'_1 \Vert _{H^1_h(\Omega )}\\&\qquad + \frac{C}{h^2} \Vert \nabla _h(u_0 - u'_0)\Vert _{H^2_h(\Omega )} \Vert \nabla _h u'_0\Vert _{H^2_h(\Omega )} \Vert \nabla _h u_2 \Vert _{L^2_h(\Omega )}\\&\qquad + \frac{|\gamma _h(u_{0,h})|}{h^2} \Vert \nabla _h(u_0 - u'_0)\Vert _{H^2_h(\Omega )} + |\gamma _h(u_{0,h}) - \gamma _h(u_{0,h}')|\\&\quad \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h}, \end{aligned}$$

where we used again Corollary 2.4, \(|P|_h = |P|\), \(|\gamma _h(u_{0,h})| \le Ch^2\) and

$$\begin{aligned} |\gamma _h(u_{0,h}) - \gamma _h(u_{0,h}')|&\le \frac{1}{h^3} \bigg \Vert \int \limits _0^1 (1-\tau ) \Big (D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h u_{0,h}]\\&\quad - D^3\tilde{W}(\nabla _h u'_{0,h})[\nabla _h u'_{0,h}, \nabla _h u'_{0,h}]\Big )\bigg \Vert _{(L^2_h)'}\\&\le CM_2 \Vert \nabla _h(u_0 - u'_0)\Vert _{H^2_h(\Omega )} \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h}. \end{aligned}$$

Finally from

$$\begin{aligned} \partial _{x_j} \mathcal {Q}_{2,h}(u_1, u_2)&= \frac{1}{h^2} \int \limits _0^1 D^3\tilde{W}(\tau \nabla _h u_0)[\nabla _h u_0, \nabla _h\partial _{x_j} u_2] d\tau + \frac{1}{h^2} D^3\tilde{W}(\nabla _h u_0) [\nabla _h \partial _{x_j} u_0, \partial _h u_2]\\&\quad + \frac{2}{h^2} D^3\tilde{W}(\nabla _h u_0) [\nabla _h \partial _{x_j} u_1, \nabla _h u_1] + \frac{1}{h^2} D^4\tilde{W}(\nabla _h u_0) [\nabla _h \partial _{x_j} u_0, \nabla _h u_1, \nabla _h u_1]\\&\quad - \frac{\gamma _h}{h^3} D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h\partial _{x_j} u_0, P] \end{aligned}$$

it follows

$$\begin{aligned} \Vert \partial _{x_j} (\mathcal {Q}_{2,h}(u_1, u_2) - \mathcal {Q}_{2,h}(u'_1, u'_2))\Vert _{(L^2_h)'} \le CM_2 \Vert (u_1 - u'_1, u_2 - u'_2)\Vert _{\mathcal {W}_h}. \end{aligned}$$

Choosing now \(M_2 \in (0,1]\) small enough we obtain that

$$\begin{aligned} \mathcal {F}_{h, f_0,f_1,f_2}:\overline{B_{CM_2 h^2}(0)} \subset \mathcal {X}_h\times \mathcal {W}_h \rightarrow \mathcal {X}_h\times \mathcal {W}_h \end{aligned}$$

defined by

$$\begin{aligned} \begin{pmatrix} u_0 \\ u_1 \\ u_2 \end{pmatrix} \mapsto \begin{pmatrix} \mathcal {G}_{h, f_0 - u_{2}} (u_0)\\ \mathcal {L}_h^{-1}\left( \begin{pmatrix} f_1 \\ f_2 \end{pmatrix}, \mathcal {Q}_{h} (u_1, u_2)\right) \end{pmatrix} \end{aligned}$$

is a \(\frac{1}{2}\)-contraction, where \(f_0 := h^2 f^h|_{t=0}\), \(f_1 := h^2\partial _t f^h|_{t=0} - u_{3,h}\) and \(f_2 := h^2\partial ^2_t f^h|_{t=0} - u_{4,h}\). We can use an analogous argument as in Lemma 3.1. First it holds, due to (3.6) and (3.5), for \(M>0\) sufficiently small

$$\begin{aligned} \Vert \mathcal {F}_{h, f_0,f_1,f_2}(0)\Vert _{\mathcal {X}_h\times \mathcal {W}_h} \le \tilde{C} M h^2 \le \frac{C M_2 h^2}{2} \end{aligned}$$

and with the \(\frac{1}{2}\)-contraction property we obtain the self mapping of \(\mathcal {F}_{h, f_0,f_1,f_2}\). Moreover, due to the norm on \(\mathcal {X}_h\) and \(\mathcal {W}_h\) we obtain (3.22) and (3.23), respectively.

Finally, the construction of \(\tilde{u}_h\) implies that \(\tilde{u}_{j,h}\) satisfies

$$\begin{aligned} \frac{1}{h^2}\Big (D^2\tilde{W}(0) \nabla _h \tilde{u}_{j,h}, \nabla _h\varphi \Big )_{L^2(\Omega )}&= \Big (h^2 \partial _t^j f^h|_{t=0} - \tilde{u}_{2+j,h}, \varphi \Big )_{L^2(\Omega )} + (\partial _t^j r_h, \varphi )_{L^2(\Omega )}\\&\quad - \frac{1}{h^2} \int \limits _0^L \Big ({\text {tr}}_{\partial S}(\partial _t^j r_{N,h}(x_1, \cdot )), {\text {tr}}_{\partial S}(\varphi (x_1, \cdot ))\Big )_{L^2(\partial S)} dx_1 \end{aligned}$$

for \(j=0,1,2\) and all \(\varphi \in \mathcal {B}\). This implies with (3.19)–(3.21)

$$\begin{aligned}&\frac{1}{h^2} \big (\varepsilon _h(u_{1,h} - \tilde{u}_{1,h}), \varepsilon _h(\varphi )\big )_{L^2(\Omega )} = -\frac{1}{h^2}\Big ((D^2\tilde{W}(\nabla _h u_{0,h})- D^2\tilde{W}(0)) \nabla _h u_{1,h}, \nabla _h \varphi \Big )_{L^2(\Omega )}\\&\quad + (r_{1,h}, \varphi )_{L^2(\Omega )} - \frac{1}{h^2} \int \limits _0^L \Big ({\text {tr}}_{\partial S}(\partial _t r_{N,h}(x_1, \cdot )), {\text {tr}}_{\partial S}(\varphi (x_1, \cdot ))\Big )_{L^2(\partial S)} dx_1\\&\frac{1}{h^2}\big (\varepsilon _h(u_{2,h} - \tilde{u}_{2,h}), \varepsilon _h(\varphi )\big )_{L^2(\Omega )} = -\frac{1}{h^2}\Big ((D^2\tilde{W}(\nabla _h u_{0,h})- D^2\tilde{W}(0)) \nabla _h u_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )}\\&\quad +(r_{2,h}, \varphi )_{L^2(\Omega )} - \frac{1}{h^2} \int \limits _0^L \Big ({\text {tr}}_{\partial S}(\partial _t^2 r_{N,h}(x_1, \cdot )), {\text {tr}}_{\partial S}(\varphi (x_1, \cdot ))\Big )_{L^2(\partial S)} dx_1\\&\quad -\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h}) [\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} - \frac{\gamma _h}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) P,\nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$

for all \(\varphi \in \mathcal {B}\), where we defined

$$\begin{aligned} r_{j,h} := u_{2+j,h} - \tilde{u}_{2+j,h} - \partial _t^j r_h. \end{aligned}$$

With this it follows \(\max _{j=1,2} \Vert r_{j,h}\Vert _{C^0(0,T;L^2(\Omega ))}\le Ch^3\) because of the definition of \(u_{2+j,h}\) and the bound on \(\partial _t r_h\). Additionally, we have due to Lemma 2.3 and Corollary 2.4, the bounds on \((u_{0,h}, u_{1,h}, u_{2,h})\) and \(\varphi \in \mathcal {B}\)

$$\begin{aligned}&\bigg | \frac{1}{h^2}\Big ((D^2\tilde{W}(\nabla _h u_{0,h})- D^2\tilde{W}(0)) \nabla _h u_{j,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} \bigg |\\&\quad = \bigg |\frac{1}{h^2} \int \limits _0^1 \Big (D^3\tilde{W}(\tau \nabla _h u_{0,h}) [\nabla _h u_{0,h}, \nabla _h u_{j,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} d\tau \bigg | \le Ch^3 \bigg \Vert \frac{1}{h}\varepsilon _h(\varphi )\bigg \Vert _{L^2(\Omega )} \end{aligned}$$

as well as

$$\begin{aligned} \bigg |\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h}) [\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )}\bigg | \le Ch^3 \bigg \Vert \frac{1}{h}\varepsilon _h(\varphi )\bigg \Vert _{L^2(\Omega )} \end{aligned}$$

and

$$\begin{aligned} \bigg |\frac{\gamma _h}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) P,\nabla _h \varphi \Big )_{L^2(\Omega )}\bigg | \le Ch^2 \bigg \Vert \frac{1}{h}\varepsilon _h(\varphi )\bigg \Vert _{L^2(\Omega )}. \end{aligned}$$

Regarding the boundary terms, we use that \({\text {tr}}_{\partial S}:H^1(S)\rightarrow H^{\frac{1}{2}}(\partial S)\) is linear and bounded. Hence, for \(j=0,1,2\)

$$\begin{aligned}&\bigg |\frac{1}{h^2} \int \limits _0^L \Big ({\text {tr}}_{\partial S}(\partial _t^j r_{N,h}(x_1, \cdot )), {\text {tr}}_{\partial S}(\varphi (x_1, \cdot ))\Big )_{L^2(\partial S)} dx_1\bigg |\nonumber \\&\quad \le \frac{1}{h^2} \Vert \partial _t^j r_{N,h}\Vert _{L^2(0,L;H^1(S))} \Vert \varphi \Vert _{L^2(0,L;H^1(S))} \le Ch^3 \bigg \Vert \frac{1}{h}\varepsilon _h(\varphi )\bigg \Vert _{L^2(\Omega )}, \end{aligned}$$
(3.30)

where we used that \(\Vert r_{N,h}\Vert _{C^2([0,T]; H^1(\Omega ))} \le Ch^5\) and the Poincaré and Korn inequality for \(\varphi \). Choosing \(\varphi = u_{j,h} - \tilde{u}_{j,h}\), it follows with an absorption argument

$$\begin{aligned} \max _{j=1,2} \bigg \Vert \frac{1}{h}\varepsilon _h(u_{j,h}) -\frac{1}{h}\varepsilon _h(\tilde{u}_{j,h})\bigg \Vert _{L^2(\Omega )} \le {\left\{ \begin{array}{ll} Ch^3,&{}\text {if}\quad j=1,\\ Ch^2,&{}\text {if}\quad j=2. \end{array}\right. } \end{aligned}$$

Now, for \(u_{0,h}-\tilde{u}_{0,h}\) it holds

$$\begin{aligned}&\frac{1}{h^2} (\varepsilon _h(u_{0,h}-\tilde{u}_{0,h}), \varepsilon _h(\varphi ))_{L^2(\Omega )} = -\frac{1}{h^2} (G(\nabla _h u_{0,h}), \nabla _h \varphi )_{L^2(\Omega )}\\&\quad + (r_{0,h}, \varphi )_{L^2(\Omega )} - \frac{1}{h^2} \int \limits _0^L \Big ({\text {tr}}_{\partial S}(r_{N,h}(x_1, \cdot )), {\text {tr}}_{\partial S}(\varphi (x_1, \cdot ))\Big )_{L^2(\partial S)} dx_1. \end{aligned}$$

The definition of G implies now

$$\begin{aligned} \bigg |\frac{1}{h^2} (G(\nabla _h u_{0,h}), \nabla _h \varphi )_{L^2(\Omega )}\bigg |&= \bigg |\frac{1}{h^2} \int \limits _0^1 (1-\tau ) \Big (D^3\tilde{W}(\tau \nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h u_{0,h}], \nabla _h\varphi \Big )_{L^2(\Omega )} d\tau \bigg |\\&\le Ch^3 \bigg \Vert \frac{1}{h}\varepsilon _h(\varphi )\bigg \Vert _{L^2(\Omega )} \end{aligned}$$

because of the bounds for \(u_{0,h}\) and Corollary 2.4. Using (3.30), it follows

$$\begin{aligned} \bigg \Vert \frac{1}{h}\varepsilon _h(u_{0,h}) -\frac{1}{h}\varepsilon _h(\tilde{u}_{0,h})\bigg \Vert _{L^2(\Omega )} \le Ch^3 \end{aligned}$$

\(\square \)

3.3 Main result

Theorem 3.3

Let \(f_h\), \(\tilde{v}_0\), \(\tilde{v}_1\), \(\tilde{u}_{j,h}\), \(j=0,1,2\) and \(\tilde{u}_h\) be given as above. Then there exists some \(h_0\in (0,1]\) such that for \(h\in (0,h_0]\) there are initial values \((u_{0,h}, u_{1,h})\) satisfying (A.1)–(A.3) and such that

$$\begin{aligned} \max _{j=0,1} \bigg \Vert \frac{1}{h} \varepsilon _h(u_{j,h}) - \frac{1}{h} \varepsilon _h(\tilde{u}_{j,h})\bigg \Vert _{L^2(\Omega )} \le Ch^3. \end{aligned}$$

Moreover, if \(u_h\) solves (3.1)–(3.4), then

$$\begin{aligned} \bigg \Vert \bigg ((u_h - \tilde{u}_h), \frac{1}{h} \int \limits _0^t \varepsilon _h\big (u_h(\tau ) - \tilde{u}_h(\tau )\big ) d\tau \bigg )\bigg \Vert _{L^\infty (0,L; L^2(\Omega ))} \le Ch^3\quad \text {for all}\quad 0<h\le h_0. \end{aligned}$$

Proof

Given \((u_{3,h}, u_{4,h})\) we construct \((u_{0,h}, u_{1,h}, u_{2,h})\) such that (A.1)–(A.3) holds. First we note that \(\Vert u_{4,h}\Vert _{L^2(\Omega )}\) is of order \(h^2\) as \(\partial _{x_1}^l v^{4}\) is bounded in \(L^2(0,L)\) for \(l=0,1\). Moreover, we have

$$\begin{aligned} \int \limits _\Omega u_{2+j, h} dx = 0 \end{aligned}$$

for \(j=1,2\) and

$$\begin{aligned} \int \limits _\Omega u_{3,h}\cdot x^\perp dx = 0. \end{aligned}$$

Using the structure of \(u_{3,h}\) we obtain

$$\begin{aligned} \frac{1}{h}\varepsilon _h(u_{3,h}) = h^2 \begin{pmatrix} -x_2 \partial _{x_1}^2 v_{2}^3 - x_3 \partial _{x_1}^2 v_{3}^3 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \end{pmatrix}. \end{aligned}$$

Altogether we obtain that \(u_{3,h}\) and \(u_{4,h}\) satisfy (A.1)–(A.3), the necessary conditions for the large times existence result in the appendix. The assumptions on g and the structure of \(f_h\) imply that (A.4) and (A.5) are fulfilled. Applying Lemma 3.1 and 3.2, we obtain for \(h_0\) sufficiently small, the existence of \((u_{0,h}, u_{1,h}, {\bar{u}}_{2,h})\) such that

$$\begin{aligned} \frac{1}{h^2} \Big (D\tilde{W}(\nabla _h u_{0,h}), \nabla _h \varphi \Big )_{L^2(\Omega )}&= (h^2 f_h|_{t=0}, \varphi )_{L^2(\Omega )} - ({\bar{u}}_{2,h}, \varphi )_{L^2(\Omega )}\\ \frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{1,h}, \nabla _h \varphi \Big )_{L^2(\Omega )}&= (h^2 \partial _t f_h|_{t=0}, \varphi )_{L^2(\Omega )} - (u_{3,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$

and

$$\begin{aligned}&\frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h {\bar{u}}_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2 \partial _t^2 f_h|_{t=0} - u_{4,h}, \varphi )_{L^2(\Omega )}\\&\quad -\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h}) [\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} - \frac{\gamma _h}{h^3} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) P,\nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$

for all \(\varphi \in \mathcal {B}\). We use the ansatz \(u_{2,h} = {\bar{u}}_{2,h} + \gamma _2^h x^\perp \) and \({\bar{u}}_{2+j,h} = u_{2+j,h} + \gamma _{2+j}^h x^\perp \) for \(j=1,2\). Choosing

$$\begin{aligned} \gamma _2^h := -\frac{1}{\mu (S) h^2} \Big (D\tilde{W}(\nabla _h u_{0,h}), \nabla _h x^\perp \Big )_{L^2} \end{aligned}$$

it follows

$$\begin{aligned} \frac{1}{h^2} \Big (D\tilde{W}(\nabla _h u_{0,h}), \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2 f_h|_{t=0}, \varphi )_{L^2(\Omega )} - (u_{2,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.31)

for all \(\varphi \in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\). Moreover, for

$$\begin{aligned} \gamma _3^h := \frac{1}{\mu (S)h^2} \Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{1,h}, \nabla _h x^\perp \Big )_{L^2} \end{aligned}$$

we deduce

$$\begin{aligned} \frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{1,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2 \partial _t f_h|_{t=0}, \varphi )_{L^2(\Omega )} - ({\bar{u}}_{3,h}, \varphi )_{L^2(\Omega )} \end{aligned}$$
(3.32)

for all \(\varphi \in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\). Then it holds \(|\gamma _2^h|\le C h^2\) as

$$\begin{aligned} D\tilde{W}(\nabla _h u_{0,h}) = D^2\tilde{W}(0)[\nabla _h u_{0,h}] + \int \limits _0^1 (1-\tau ) D^3\tilde{W}(\tau \nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h u_{0,h}] d\tau \end{aligned}$$

and \(|\gamma _3^h| \le C h^2\) with a similar calculation. Lastly, we need to find \(\gamma _4^h\) such that

$$\begin{aligned}&\frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{2,h}, \nabla _h \varphi \Big )_{L^2(\Omega )} = (h^2 \partial _t^2 f_h|_{t=0} - {\bar{u}}_{4,h}, \varphi )_{L^2(\Omega )}\nonumber \\&\quad -\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h}) [\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h \varphi \Big )_{L^2(\Omega )} \end{aligned}$$
(3.33)

for all \(\varphi \in H^1_{\textrm{per}}(\Omega ;\mathbb {R}^3)\). Therefore, we choose

$$\begin{aligned} \gamma _4^h&:= -\frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h u_{2,h}, \nabla _h x^\perp \Big )_{L^2} -\frac{\gamma _2^h}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h x^\perp , \nabla _h x^\perp \Big )_{L^2}\\&\quad + \frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h x^\perp \Big )_{L^2}. \end{aligned}$$

The first and last term can be bounded easily, using Corollary 2.4

$$\begin{aligned} \bigg |\frac{1}{h^2}\Big (D^2\tilde{W}(\nabla _h u_{0,h}) \nabla _h {\bar{u}}_{0,h}, \nabla _h x^\perp \Big )_{L^2} \bigg |&= \bigg |\frac{1}{h^3} \int \limits _0^1 \Big (D^3\tilde{W}(\nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h{\bar{u}}_{2,h}], P\Big )_{L^2}\bigg | \\&\le \frac{C}{h^2} \Vert \nabla _h u_{0,h}\Vert _{H^1_h} \Vert \nabla _h {\bar{u}}_{2,h}\Vert _{H^1_h} \le Ch^2 \end{aligned}$$

and

$$\begin{aligned} \bigg |\frac{1}{h^2} \Big (D^3\tilde{W}(\nabla _h u_{2,h})[\nabla _h u_{1,h}, \nabla _h u_{1,h}], \nabla _h x^\perp \Big )_{L^2} \bigg | \le Ch^2. \end{aligned}$$

For the second part of \(\gamma _4^h\), we use the following equality

$$\begin{aligned} \Big (D^2\tilde{W}(\nabla _h u_{0,h}) P, P\Big )_{L^2(\Omega )}&= \Big (D^3\tilde{W}(0) [\nabla _h u_{0,h}, P], P\Big )_{L^2(\Omega )} \\&\quad + \int \limits _0^1 (1-\tau ) \Big (D^4\tilde{W}(\tau \nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h u_{0,h}, P], P\Big )_{L^2(\Omega )} d\tau \end{aligned}$$

where

$$\begin{aligned} \bigg |\Big (D^4\tilde{W}(\tau \nabla _h u_{0,h})[\nabla _h u_{0,h}, \nabla _h u_{0,h}, P], P\Big )_{L^2(\Omega )}\bigg | \le Ch^4 \quad \text {for all}\quad \tau \in [0,1] \end{aligned}$$
(3.34)

as \(\Vert u_{0,h}\Vert _{H^1_h(\Omega )} \le Ch^2\) and \(|P|_h = |P|\), because \(P\in \mathbb {R}^{3\times 3}_{skew}\). Furthermore, we obtain with (2.5)

$$\begin{aligned} \Big (D^3\tilde{W}(0) [\nabla _h u_{0,h}, P], P\Big )_{L^2(\Omega )}&= h \bigg (D^3\tilde{W}(0) \bigg [\frac{1}{h}\varepsilon _h(u_{0,h}) - \frac{1}{h}\varepsilon _h(\tilde{u}_{0,h}), P\bigg ], P\bigg )_{L^2(\Omega )}\\&\quad + \Big (D^3\tilde{W}(0) [\nabla _h \tilde{u}_{0,h}, P], P\Big )_{L^2(\Omega )}. \end{aligned}$$

Utilizing the inequality for the initial values (3.24), we deduce

$$\begin{aligned} \bigg | h \bigg (D^3\tilde{W}(0) \bigg [\frac{1}{h}\varepsilon _h(u_{0,h}) - \frac{1}{h}\varepsilon _h(\tilde{u}_{0,h}), P\bigg ], P\bigg )_{L^2}\bigg | \le C h^4. \end{aligned}$$

Lastly due to the symmetry properties of \(D^3\tilde{W}\), the structure of \(\nabla _h \tilde{u}_{0,h}\) and (2.5), it follows

$$\begin{aligned} \bigg |\Big (D^3\tilde{W}(0) [{\text {sym}}(\nabla _h \tilde{u}_{0,h}), P], P\Big )_{L^2}\bigg | = \bigg |\Big (D^3\tilde{W}(0) [h^3 Q, P], P\Big )_{L^2} + \Big (D^3\tilde{W}(0) [R, P], P\Big )_{L^2}\bigg | \end{aligned}$$

where

$$\begin{aligned} Q&= \begin{pmatrix} -x_2 \partial _{x_1}^2 v_2 - x_3 \partial _{x_1}^2 v_3 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \end{pmatrix}\\ R&= h^4 {\text {sym}}\begin{pmatrix} 0 &{}\quad \partial _{x_2} a_2 \partial _{x_1}^3 v_2 + \partial _{x_2} a_3 \partial _{x_1}^3 v_3 &{}\quad \partial _{x_3} a_2 \partial _{x_1}^3 v_2 + \partial _{x_3} a_3\partial _{x_1}^3 v_3 \\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\end{pmatrix} \\&\quad +h^5 {\text {sym}}\begin{pmatrix} a_2 \partial _{x_1}^4 v_2 + a_3 \partial _{x_1}^4 v_3 &{}\quad 0 &{}\quad 0\\ 0&{}\quad \partial _{x_2} b_2 \partial _{x_1}^4 v_2 + \partial _{x_2} c_3 \partial _{x_1}^4 v_3 &{}\quad \partial _{x_3} b_2 \partial _{x_1}^4 v_2 +\partial _{x_3} c_3 \partial _{x_1}^4 v_3\\ 0 &{}\quad \partial _{x_2} b_3 \partial _{x_1}^4 v_3 +\partial _{x_2} c_2 \partial _{x_1}^4 v_2 &{}\quad \partial _{x_3} b_3 \partial _{x_1}^4 v_3 + \partial _{x_3} c_2 \partial _{x_1}^4 v_2 \end{pmatrix}\\&\quad + h^6 {\text {sym}}\begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ b_2 \partial _{x_1}^5 v_2 + c_3 \partial _{x_1}^5 v_3 &{}\quad 0 &{}\quad 0 \\ b_3 \partial _{x_1}^5 v_3 + c_2 \partial _{x_1}^5 v_2 &{}\quad 0 &{}\quad 0 \end{pmatrix}. \end{aligned}$$

Due to the structure of Q and \(P = \nabla x^\perp \), it follows

$$\begin{aligned} D^3\tilde{W}(0)[Q,P,P] = \Big ((Q^T- Q)^T {\text {sym}}(P) + (P^T - P)^T {\text {sym}}(Q)\Big ) : P = 0. \end{aligned}$$

Hence, with \(R = O(h^4)\) we obtain

$$\begin{aligned} \bigg |\Big (D^3\tilde{W}(0) [{\text {sym}}(\nabla _h \tilde{u}_{0,h}), P], P\Big )_{L^2}\bigg | \le Ch^4. \end{aligned}$$

Thus, altogether, it follows with \(|\gamma _2^h| \le Ch^2\)

$$\begin{aligned} \bigg |\frac{\gamma _2^h}{h^2} \Big (D^2\tilde{W}(\nabla _h u_{0,h})\nabla _h x^\perp , \nabla _h x^\perp \Big )_{L^2}\bigg | \le Ch^2. \end{aligned}$$

We obtain for \(h_0\) sufficiently small, the existence of \((u_{0,h}, u_{1,h}, u_{2,h})\) such that (A.1)–(A.3) are satisfied and

$$\begin{aligned} \max _{j=0,1} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(u_{j,h}) - \frac{1}{h}\varepsilon _h(\tilde{u}_{j,h})\bigg )\bigg \Vert _{L^2(\Omega )} \le Ch^3 \end{aligned}$$

holds.

Due to Theorem A.1, there exists a solution \(u_h\in \bigcap _{k=0}^4 C^k([0,T]; H^{4-k}_{\textrm{per}}(\Omega ;\mathbb {R}^3))\) of (3.1)–(3.4). Thus, \(w_h := u_h - \tilde{u}_h\) solves the system

$$\begin{aligned}&-\big (\partial _t w_h, \partial _t\varphi \big )_{L^2(Q_T)} + \frac{1}{h^2} \big (D^2\tilde{W}(0) \nabla _h w_h, \nabla _h \varphi \big )_{L^2(Q_T)} - (w_1, \varphi |_{t=0})_{L^2(\Omega )}\\&\quad = \frac{1}{h^2}\int \limits _0^1 \big ((D^2\tilde{W}(\tau \nabla _h u_h) - D^2\tilde{W}(0))\nabla _h\tilde{u}_h, \nabla _h \varphi \big )_{L^2(Q_T)}d\tau - \big (r_h, \varphi \big )_{L^2(Q_T)}\\&\qquad - \frac{1}{h^2} ({\text {tr}}_{\partial \Omega }(r_{N,h}), {\text {tr}}_{\partial \Omega }(\varphi ))_{L^2(0,T;L^2(\partial \Omega ))},\\&\quad w_h \text { is } L\text {-periodic in } x_1\text {-direction},\\&\quad w_h|_{t=0} = w_{0,h} \end{aligned}$$

for all \(\varphi \in C^1([0,T]; H^1_{\textrm{per}, (0)}(\Omega ;\mathbb {R}^3))\) with \(\varphi |_{t=T} = 0\) and with \(w_{j,h} := u_{j, h} - \tilde{u}_{j,h}\), \(j=0,1\). Hence, with (A.16) we obtain an upper bound for w. For this we use that, due to the structure of \(r_h\) and \(r_{N,h}\), it follows

$$\begin{aligned} \frac{1}{h^2} \Vert r_{N,h}\Vert _{L^1(0,T;H^1)} \le Ch^3,\quad \Vert r_h\Vert _{L^1(0,T;L^2)} \le Ch^3 \end{aligned}$$

as a, b, c and v are sufficiently regular. Moreover, using (3.24)

$$\begin{aligned} \Vert w_{k,h}\Vert _{L^2(\Omega )} \le \max _{j=0,1} \bigg \Vert \bigg (\frac{1}{h}\varepsilon _h(u_{j,h}) - \frac{1}{h}\varepsilon _h(\tilde{u}_{j,h})\bigg )\bigg \Vert _{L^2(\Omega )} \le Ch^3 \end{aligned}$$

for \(k=0,1\), where we used Poincaré’s and Korn’s inequality, as well as the fact that \(w_{k,h}\in \mathcal {B}\) holds for \(k=0,1\). With the fundamental theorem of calculus and Corollary 2.4, we deduce

$$\begin{aligned}&\sup _{\varphi \in X_h, \Vert \varphi \Vert _{X_h} = 1} \bigg |\frac{1}{h^2}\int \limits _0^1\Big ((D^2\tilde{W}(\tau \nabla _h u_h) - D^2\tilde{W}(0)) \nabla _h \partial _t \tilde{u}_h, \nabla _h \varphi \Big )_{L^2(\Omega )}d\tau \bigg |\\&\quad \le \sup _{\varphi \in X_h, \Vert \varphi \Vert _{X_h} = 1} \bigg |\frac{1}{h^2}\int \limits _0^1 \int \limits _0^1 \Big (D^3\tilde{W}(s \tau \nabla _h u_h) [\nabla _h u_h, \nabla _h \partial _t \tilde{u}_h], \nabla _h \varphi \Big )_{L^2(\Omega )}ds d\tau \bigg |\\&\quad \le \sup _{\varphi \in X_h, \Vert \varphi \Vert _{X_h} = 1} \frac{C}{h}\Vert \nabla _h u_h\Vert _{H^2_h(\Omega )}\Vert \nabla _h \partial _t \tilde{u}_h \Vert _{L^2_h(\Omega )}\Vert \nabla _h \varphi \Vert _{L^2_h(\Omega )} \le CRh^3. \end{aligned}$$

Lastly, we have to deal with the rotational term. Using the momentum balance law, \(u_{0,h}\), \(u_{1,h}\in \mathcal {B}\) and the structure of g, we obtain with \(q^h := h^2 f^h\)

$$\begin{aligned} \int \limits _0^t \int \limits _\Omega u_h\cdot x^\perp \textrm{d}x\textrm{d}\tau&= t \int \limits _\Omega u_{0,h} \cdot x^\perp dx + \frac{1}{2} t^2\int \limits _\Omega u_{1,h}\cdot x^\perp \textrm{d}x + \int \limits _0^t (t-s) \int \limits _\Omega q^h \cdot x^\perp \textrm{d}x\textrm{d}s \\&\quad +\frac{1}{h} \int \limits _0^t \int \limits _0^\tau (\tau -s) \int \limits _\Omega q^h \cdot u_h^\perp - u_h^\perp \cdot \partial _t^2 u_h \textrm{d}x\textrm{d}s\textrm{d}\tau \\&= \frac{1}{h} \int \limits _0^t \int \limits _0^\tau (\tau -s) \int \limits _\Omega q^h \cdot u_h^\perp - u_h^\perp \cdot \partial _t^2 u_h \textrm{d}x\textrm{d}s\textrm{d}\tau . \end{aligned}$$

Hence, it follows

$$\begin{aligned} \bigg \Vert \int \limits _0^t \frac{1}{h} \int \limits _\Omega u_h\cdot x^\perp \textrm{d}x\textrm{d}\tau \bigg \Vert _{C^0([0,T])}&\le C \bigg (\bigg \Vert \frac{1}{h^2}\int \limits _\Omega q^h \cdot u_h^\perp \textrm{d}x\bigg \Vert _{C^0([0,T])} + \bigg \Vert \frac{1}{h^2}\int \limits _\Omega \partial _t^2 u_h \cdot u_h^\perp \textrm{d}x\bigg \Vert _{C^0([0,T])} \bigg )\\&\le Ch^{3} \end{aligned}$$

as due to (A.6)

$$\begin{aligned} \bigg \Vert \frac{1}{h}\varepsilon (\partial _t^\delta u_h)\bigg \Vert _{L^\infty (0,T;L^2)} + \bigg \Vert \frac{1}{h}\int \limits _\Omega \partial _t^\delta u_h\cdot x^\perp \textrm{d}x\bigg \Vert _{L^\infty (0,T)} \le Ch^{2} \end{aligned}$$

for \(\delta = 0,2\). Thus, with (A.16) it follows

$$\begin{aligned} \bigg \Vert \bigg ((u_h - \tilde{u}_h), \frac{1}{h} \int \limits _0^t \varepsilon _h\big (u_h(\tau ) - \tilde{u}_h(\tau )\big ) \textrm{d}\tau \bigg )\bigg \Vert _{C^0([0,T];L^2)} \le Ch^{3}. \end{aligned}$$

\(\square \)