This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Phase field topology optimisation for 4D printing

Harald Garcke 111Fakultät für Mathematik, Universität Regensburg, 93040 Regensburg, Germany (H[email protected]).    Kei Fong Lam 222Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong (a[email protected]).    Robert Nürnberg 333Department of Mathematics, University of Trento, Trento, Italy (r[email protected]).    Andrea Signori 444Department of Mathematics, Politecnico di Milano, 20133 Milano, Italy (a[email protected]).
Abstract

This work concerns a structural topology optimisation problem for 4D printing based on the phase field approach. The concept of 4D printing as a targeted evolution of 3D printed structures can be realised in a two-step process. One first fabricates a 3D object with multi-material active composites and apply external loads in the programming stage. Then, a change in an environmental stimulus and the removal of loads cause the object deform in the programmed stage. The dynamic transition between the original and deformed shapes is achieved with appropriate applications of the stimulus. The mathematical interest is to find an optimal distribution for the materials such that the 3D printed object achieves a targeted configuration in the programmed stage as best as possible.

Casting the problem as a PDE-constrained minimisation problem, we consider a vector-valued order parameter representing the volume fractions of the different materials in the composite as a control variable. We prove the existence of optimal designs and formulate first order necessary conditions for minimisers. Moreover, by suitable asymptotic techniques, we relate our approach to a sharp interface description. Finally, the theoretical results are validated by several numerical simulations both in two and three space dimensions.

Keywords: 4D printing, Printed active composites, Topology optimisation, Phase field, Linear elasticity, Optimal control AMS (MOS) Subject Classification: 49J20, 49K40, 49J50

1 Introduction

Four-dimensional (4D) printing [2, 52, 61] entails the combination of additive manufacturing (3D printing) and active material technologies to create printed composites capable of morphing into different configurations in response to various environmental stimuli. First designs of such composites consist of active material components, such as piezoelectric ceramics, hydrogels or shape memory polymers [24], in the form of fibers integrated within a passive elastomeric matrix [37]. These multi-material active composites were originally difficult to manufacture, owing to the fragility of the materials involved [44]. However, with 3D printing techniques it is nowadays feasible to fabricate these active composites to a high degree of precision, resulting in so-called printed active composites (PACs) [50]. For an overview of other 4D printing strategies besides PACs in the construction of smart materials allowing direct stimuli-responsive transformations, we refer to [67].

The shape shifting functionality of the active components enables the self-actuating and self-assembling potentials of PACs, allowing them to fold, bend, twist, expand and contract when a stimulus is applied, and return to their original configurations after the stimulus is removed. This property has led to the fabrication of intelligent active hinges and origami-like objects [35, 37], mesh structures [24, 64] and self-actuated deformable solids [59] in the form of in the form of graspers and smart key-lock systems. We refer to the review article [52] and the references cited therein for more applications of 4D printing. The shape memory behaviour of the PACs can be programmed in a two-step cycle: The first (programming) step involves deforming the structure from its permanent shape to a metastable temporary shape, and the second (recovery) step involves applying an appropriate stimulus so that the structure regains its original shape. A typical stimulus is heat (in combination with light [36] or water [7]), in which programmed PACs alter their shapes when the temperature rises above or drops below a critical value.

With the advances in the state-of-the art 3D printing technologies, the designs of PACs need not be limited to the conventional fibre-matrix architectures first considered in [37]. In particular, the distribution of active and passive materials in the designs can take on more complicated geometries to better fulfil the intended functionalities of the PACs. This opens up the possibility of a computational design approach guided by a structural topology optimisation framework. In the context of 3D printing, see, e.g., the review [26], this framework has been applied to explore optimising support structures to overhang regions [3, 46, 51], as well as self-support designs respecting the overhang angle constraints [20, 32, 43, 45]. For active materials and active composites, [40, 54] studied how to pattern thin-film layers within a multi-layer structure with the aim of generating large shape changes via spatially varying eigenstrains within the microstructures, while [50] aimed to optimise the microstructures of PACs matching various target shapes after a thermomechanical training and activation cycle. Later works incorporated nonlinear thermoelasticity [39, 59], thermo-mechanical cycles of shape memory polymers [12], reversible deformations [49], as well as multi-material designs [65] within the topology optimisation framework.

In many of the aforementioned contributions, the topology optimisation is implemented numerically with the level-set method or the solid isotropic material with penalisation (SIMP) approach. In this work we employ an alternative approach based on the phase field methodology [17], which allows a straightforward extension to the multiphase setting [13, 62] involving multiple (possibly distinct) types of active materials within the design. In particular, this opens up the design to multiphase PACs that can memorise more than two shapes [38, 47, 58, 63, 66]. The phase field-based structural topology optimisation approach has been popularised in recent years by many authors, with applications in nonlinear elasticity [55], stress constraints [19], compliance optimisation [15, 60], elastoplasticity [4], eigenfrequency maximisation [31, 60], graded-material design [21], shape optimisation in fluid flow [28, 29, 30] and more recently for 3D printing with overhang angle constraints [32].

Taking inspiration from the setting of Maute et al. [50], we formulate a structural topology optimisation problem for a multiphase PAC with the objective of finding optimal distributions of active and passive materials so that the resulting composite matches targeted shapes as close as possible. An additional perimeter regularisation term, in the form of a multiphase Ginzburg–Landau functional, is added, and our contribution involves a mathematical analysis of the resulting multiphase structural topology optimisation problem with emphasis on the rigorous derivation of minimisers and optimality conditions. A sharp interface asymptotic analysis is performed to obtain a set of optimality conditions applicable in a level set-based shape optimisation framework. We perform numerical simulations in two and three spatial dimensions to show the optimal distributions of active and passive components in order to match with various target shapes for the PACs.

The rest of this paper is organised as follows: in Section 2 we formulate the phase field structural optimisation problem to be studied, and present several preliminary mathematical results. In Sections 3 and 4 we analyse the design optimisation problem and establish analytical results concerning minimisers and optimality conditions. The sharp interface limit is explored in Section 5 and, finally, in Section 6 we present the numerical discretisation and several simulations of our approach.

2 Problem formulation

Within a bounded domain Ωd\Omega\subset{\mathbb{R}}^{d}, d{2,3}d\in\{2,3\}, with Lipschitz boundary Γ:=Ω\Gamma:=\partial\Omega, we assume there are LL types of linearly elastic materials, whose volume fractions are encoded with the help of a vectorial phase field variable 𝝋=(φ1,,φL):ΩΔL{\bm{\varphi}}=(\varphi_{1},\dots,\varphi_{L}):\Omega\to\Delta^{L}, where ΔL\Delta^{L} denotes the Gibbs simplex in L{\mathbb{R}}^{L}:

ΔL:={𝒙=(x1,,xL)L:i=1Lxi=1,xi0 for all i{1,,L}}.\Delta^{L}:=\Big{\{}\bm{x}=(x_{1},...,x_{L})\in{\mathbb{R}}^{L}\,:\,\sum_{i=1}^{L}x_{i}=1,\,x_{i}\geq 0\text{ for all }i\in\{1,\dots,L\}\Big{\}}.

For our application to PACs, we take φL\varphi_{L} as the volume fraction of the passive elastic material, and φ1,,φL1\varphi_{1},\dots,\varphi_{L-1} as the volume fractions of (possibly different) active elastic materials. Note that in the two-phase case L=2L=2, we simply have 𝝋=(φ1,φ2){\bm{\varphi}}=(\varphi_{1},\varphi_{2}), and due to the relation φ1+φ2=1\varphi_{1}+\varphi_{2}=1 we may instead use the scalar difference function φ:=φ1φ2\varphi:=\varphi_{1}-\varphi_{2} to encode 𝝋{\bm{\varphi}} via the relation 𝝋=(12(1+φ),12(1φ)){\bm{\varphi}}=(\frac{1}{2}(1+\varphi),\frac{1}{2}(1-\varphi)). This particular scenario will be employed later on, when dealing with the connection between the problem we are going to analyse and the corresponding sharp interface limit in Section 5, as well as for the numerical simulations presented in Section 6.

The shape shifting mechanism considered in [50, 68] involves two levels of temperature and one set of external loads, with one temperature THT_{H} higher than a critical transition temperature TgT_{g} of the active materials (e.g., the glass transition temperature for shape memory polymers), and the other temperature TLT_{L} lower than the critical temperature. The printed composite is first heated to THT_{H}, and the shape memory cycle starts at THT_{H} and proceeds as follows: First, external loads are applied to deform the printed composite while the temperature remains at THT_{H}, with the new configuration being known as the programming stage (or Stage 1). Then, the temperature is decreased while the loads are maintained on the printed composite, which are then removed once the temperature reached TLT_{L}. The resulting shape at TLT_{L} is the desired shape and we denote it as the programmed stage (or Stage 2). Increasing the temperature to THT_{H} enables the printed composite to recover its original shape, and this ends the shape memory cycle, see Figure 1 for the thermo-mechanical processing steps involving the two stages.

Refer to caption
Figure 1: Schematics of the the shape memory cycle from [50] involving a programming stage (Stage 1) and a programmed stage (Stage 2).

To capture the above behaviour, following [50] we consider a model for each stage. In the programming stage (Stage 1), we consider an elastic displacement 𝐮¯:Ωd{\overline{{\bf u}}}:\Omega\to{\mathbb{R}}^{d} and decompose the domain boundary Γ\Gamma into a partition Γ=cl(Γ¯D)cl(Γ¯N)\Gamma={\rm cl}(\overline{\Gamma}_{D})\cup{\rm cl}(\overline{\Gamma}_{N}) with relative open subsets Γ¯D\overline{\Gamma}_{D} and Γ¯N\overline{\Gamma}_{N} such that Γ¯DΓ¯N=\overline{\Gamma}_{D}\cap\overline{\Gamma}_{N}=\emptyset and Γ¯D\overline{\Gamma}_{D}\neq\emptyset, where cl(A){\rm cl}(A) denotes the closure of a set AA, and we assign a prescribed displacement 𝑼¯\overline{\bm{U}} on Γ¯D\overline{\Gamma}_{D} and surface loads 𝐠¯\overline{{\bf g}} on Γ¯N\overline{\Gamma}_{N}. Under a linearised elasticity setting, the balance of momentum yields the following system of equations for the displacement 𝐮¯{\overline{{\bf u}}}:

div(¯(𝝋)(𝐮¯))\displaystyle-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})\big{)} =𝐅¯\displaystyle=\overline{{\bf F}}  in Ω,\displaystyle\quad\text{ in }\Omega, (2.1a)
𝐮¯\displaystyle{\overline{{\bf u}}} =𝑼¯\displaystyle=\overline{\bm{U}}  on Γ¯D,\displaystyle\quad\text{ on }\overline{\Gamma}_{D}, (2.1b)
(¯(𝝋)(𝐮¯))𝐧\displaystyle\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})\big{)}{\bf n} =𝐠¯\displaystyle=\overline{{\bf g}}  on Γ¯N,\displaystyle\quad\text{ on }\overline{\Gamma}_{N}, (2.1c)

with a phase-dependent elasticity tensor ¯\overline{{\mathbb{C}}}, body force 𝐅¯\overline{{\bf F}}, outer unit normal 𝐧{\bf n}, and symmetrised gradient (𝐮¯)\mathcal{E}({\overline{{\bf u}}}). One example of ¯(𝝋)\overline{{\mathbb{C}}}({\bm{\varphi}}) is

¯(𝝋(𝒙))=i=1L¯iφi(𝒙) for 𝝋(𝒙)=(φ1(𝒙),,φL(𝒙))ΔL,𝒙Ω,\overline{{\mathbb{C}}}({\bm{\varphi}}(\bm{x}))=\sum_{i=1}^{L}\overline{{\mathbb{C}}}_{i}\varphi_{i}(\bm{x})\quad\text{ for }{\bm{\varphi}}(\bm{x})=(\varphi_{1}(\bm{x}),\dots,\varphi_{L}(\bm{x}))\in\Delta^{L},\quad\bm{x}\in\Omega,

with constant tensors ¯i\overline{{\mathbb{C}}}_{i}, 1iL1\leq i\leq L.

After the change in temperature from THT_{H} to TLT_{L} and after the programming loads in Stage 1 have been removed, the PAC experiences deformations due to residual stresses generated during the thermomechanical processing steps. When the temperature falls below TgT_{g}, the active elastic materials undergo a phase transition from a soft rubbery state to a glassy state that has a higher Young’s modulus. We introduce a new variable 𝐮^:Ωd{\widehat{{\bf u}}}:\Omega\to{\mathbb{R}}^{d} to denote the displacement in the programmed stage (Stage 2), and as in [50], model the strains from the programming stage (Stage 1) as eigenstrains for 𝐮^{\widehat{{\bf u}}}. These eigenstrains are present only in the regions of active elastic materials, which we model with a fixity function χ:L[0,)\chi:{\mathbb{R}}^{L}\to[0,\infty). The shape fixity for a shape memory material is the ratio (expressed as a percentage) between the strain in the stress-free state after the programming step and the maximum strain [1]. For example, if the deformation elongates the material, the fixity quantifies the ability of the material to hold the temporary elongated length when the stress is removed. It is clear from the definition that for a passive elastic material, the fixity is zero, and so we set that χ=0\chi=0 in the region {φL=1}\{\varphi_{L}=1\} of the passive elastic material. Decomposing the domain boundary Γ\Gamma into a possibly different partition Γ=cl(Γ^D)cl(Γ^N)\Gamma={\rm cl}(\widehat{\Gamma}_{D})\cup{\rm cl}(\widehat{\Gamma}_{N}) with relative open subsets Γ^D\widehat{\Gamma}_{D} and Γ^N\widehat{\Gamma}_{N} such that Γ^DΓ^N=\widehat{\Gamma}_{D}\cap\widehat{\Gamma}_{N}=\emptyset and Γ^D\widehat{\Gamma}_{D}\neq\emptyset, where we assign a prescribed displacement 𝑼^\widehat{\bm{U}} on Γ^D\widehat{\Gamma}_{D} and surface loads 𝐠^\widehat{{\bf g}} on Γ^N\widehat{\Gamma}_{N}, the equations for the programmed stage (Stage 2) read as

div(^(𝝋)((𝐮^)χ(𝝋)(𝐮¯)))\displaystyle-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}))\big{)} =𝐅^\displaystyle=\widehat{{\bf F}} in Ω,\displaystyle\text{ in }\Omega, (2.2a)
𝐮^\displaystyle{\widehat{{\bf u}}} =𝑼^\displaystyle=\widehat{\bm{U}} on Γ^D,\displaystyle\text{ on }\widehat{\Gamma}_{D}, (2.2b)
(^(𝝋)((𝐮^)χ(𝝋)(𝐮¯)))𝐧\displaystyle\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}))\big{)}{\bf n} =𝐠^\displaystyle=\widehat{{\bf g}} on Γ^N,\displaystyle\text{ on }\widehat{\Gamma}_{N}, (2.2c)

with a phase-dependent elasticity tensor ^\widehat{{\mathbb{C}}} and body force 𝐅^\widehat{{\bf F}}. In the above, the change in the elasticity tensor from ¯\overline{{\mathbb{C}}} in Stage 1 to ^\widehat{{\mathbb{C}}} in Stage 2 encodes the change in the elastic properties of the PAC when the temperature changes from THT_{H} to TLT_{L}. Similarly to [50], here we have neglected the strains arising from thermal expansion in (2.1) and (2.2).

In the next section, under a suitable functional framework, we demonstrate that (2.1) and (2.2) are uniquely solvable, with the solution depending continuously on 𝝋{\bm{\varphi}}. Since 𝝋{\bm{\varphi}} controls the distribution of the passive and active elastic materials, it is natural to ask for specific material distributions that optimise certain cost functionals related to the design of PACs. Motivated from [50], we primarily focus on the following cost functional

J(𝝋,𝐮^):=12Γtar(W(𝐮^𝒖tar))(𝐮^𝒖tar)dd1+γΩε|𝝋|2+1εΨ(𝝋)dx,\displaystyle J({\bm{\varphi}},{\widehat{{\bf u}}}):=\frac{1}{2}\int_{\Gamma^{\rm tar}}\Big{(}W({\widehat{{\bf u}}}-\bm{u}^{\rm tar})\Big{)}\cdot({\widehat{{\bf u}}}-\bm{u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1}+\gamma\int_{\Omega}\varepsilon|\nabla{\bm{\varphi}}|^{2}+\frac{1}{\varepsilon}\Psi({\bm{\varphi}})\,\mathrm{dx}, (2.3)

where γ>0\gamma>0 is a weighting factor, 𝐮^{\widehat{{\bf u}}} is a solution to (2.2) depending on 𝝋{\bm{\varphi}} (and also on 𝐮¯{\overline{{\bf u}}}, a solution to (2.1)), Wd×dW\in{\mathbb{R}}^{d\times d} is a fixed weighting matrix, Γtar\Gamma^{\rm tar} is a subset of the boundary Γ^N\widehat{\Gamma}_{N}, ε>0\varepsilon>0 is a fixed constant related to the thickness of the interfacial regions {0<φi<1}\{0<\varphi_{i}<1\}, i{1,,L}i\in\{1,...,L\}, d1\mathcal{H}^{d-1} indicates the standard (d1)(d-1)-dimensional Hausdorff measure, and Ψ:L\Psi:{\mathbb{R}}^{L}\to{\mathbb{R}} is a non-negative multi-well potential that attains its minimum at the corners {𝒆1,,𝒆L}\{\bm{e}_{1},\dots,\bm{e}_{L}\} (the unit vectors in L{\mathbb{R}}^{L}) of the Gibbs simplex ΔL\Delta^{L}. The first term in (2.3) consists of a target shape matching term, where we like to match the displacement 𝐮^{\widehat{{\bf u}}} in Stage 2 with a prescribed deformation 𝐮tar{\bf u}^{\rm tar} over the surface ΓtarΓ^N\Gamma^{\rm tar}\subset\widehat{\Gamma}_{N} by minimising the squared difference weighted by a matrix WW. The second term is the well-known Ginzburg–Landau functional in the multiphase setting that serves as a form of perimeter regularisation. It provides some form of regularity to our design solutions and penalises designs that have large interfaces between the different phases of elastic materials.

For our problem we introduce the design space

𝒰ad={𝝋H1(Ω,L):𝝋(𝒙)ΔL for a.e. 𝒙Ω},\mathcal{U}_{\rm ad}=\Big{\{}{\bm{\varphi}}\in H^{1}(\Omega,{\mathbb{R}}^{L})\,:\,{\bm{\varphi}}(\bm{x})\in\Delta^{L}\text{ for a.e.~{}}\bm{x}\in\Omega\Big{\}},

and our design problem can be formulated as the following

(𝐏)\displaystyle{\bf(P)}\quad min𝝋𝒰adJ(𝝋,𝐮^) subject to (𝝋,𝐮¯,𝐮^) is a solution to (2.1)(2.2).\displaystyle\min_{{\bm{\varphi}}\in\mathcal{U}_{\rm ad}}J({\bm{\varphi}},{\widehat{{\bf u}}})\text{ subject to }({\bm{\varphi}},{\overline{{\bf u}}},{\widehat{{\bf u}}})\text{ is a solution to }\eqref{bu:sys}-\eqref{hu:sys}.
Remark 2.1.

For the existence theory for optimal designs to (𝐏)(\bf P), it is also possible to consider a more general form of the cost functional:

J(𝝋,𝐮¯,𝐮^)\displaystyle J({\bm{\varphi}},{\overline{{\bf u}}},{\widehat{{\bf u}}}) =ΩhΩ(𝒙,𝐮¯,𝐮^)dx+Γ¯Nh¯(𝒔,𝐮¯)dd1+Γ^Nh^(𝒔,𝐮^)dd1\displaystyle=\int_{\Omega}h_{\Omega}(\bm{x},{\overline{{\bf u}}},{\widehat{{\bf u}}})\,\mathrm{dx}+\int_{\overline{\Gamma}_{N}}\overline{h}({\bm{s}},{\overline{{\bf u}}})\,\mathrm{d}\mathcal{H}^{d-1}+\int_{\widehat{\Gamma}_{N}}\widehat{h}(\bm{s},{\widehat{{\bf u}}})\,\mathrm{d}\mathcal{H}^{d-1}
+γΩε|𝝋|2+1εΨ(𝝋)dx,\displaystyle\quad+\gamma\int_{\Omega}\varepsilon|\nabla{\bm{\varphi}}|^{2}+\frac{1}{\varepsilon}{\Psi({\bm{\varphi}})}\,\mathrm{dx},

with Carathéodory functions hΩh_{\Omega}, h¯\overline{h} and h^\widehat{h} satisfying (see [14, Remark 5])

|hΩ(𝒙,𝒗,𝒘)|\displaystyle|h_{\Omega}(\bm{x},\bm{v},\bm{w})| a1(𝒙)+b1(𝒙)|𝒗|p+b2(𝒙)|𝒘|p\displaystyle\leq a_{1}(\bm{x})+b_{1}(\bm{x})|\bm{v}|^{p}+b_{2}(\bm{x})|\bm{w}|^{p}\quad for all 𝒗,𝒘d, a.e. 𝒙Ω,\displaystyle\text{ for all }\bm{v},\bm{w}\in{\mathbb{R}}^{d},\text{ a.e.~{}}\bm{x}\in\Omega,
|h¯(𝒔,𝒗)|\displaystyle|\overline{h}(\bm{s},\bm{v})| a2(𝒔)+c1(𝒔)|𝒗|2\displaystyle\leq a_{2}(\bm{s})+c_{1}(\bm{s})|\bm{v}|^{2}\quad for all 𝒗d, a.e. 𝒔Γ¯N,\displaystyle\text{ for all }\bm{v}\in{\mathbb{R}}^{d},\text{ a.e.~{}}\bm{s}\in\overline{\Gamma}_{N},
|h¯(𝒔,𝒘)|\displaystyle|\overline{h}(\bm{s},\bm{w})| a3(𝒔)+c2(𝒔)|𝒘|2\displaystyle\leq a_{3}(\bm{s})+c_{2}(\bm{s})|\bm{w}|^{2}\quad for all 𝒘d, a.e. 𝒔Γ^N,\displaystyle\text{ for all }\bm{w}\in{\mathbb{R}}^{d},\text{ a.e.~{}}\bm{s}\in\widehat{\Gamma}_{N},

for any 2p<2\leq p<\infty if d=2d=2 and 2p<62\leq p<6 if d=3d=3, with functions a1L1(Ω)a_{1}\in L^{1}(\Omega), b1,b2L(Ω)b_{1},b_{2}\in L^{\infty}(\Omega), a2L1(Γ¯N)a_{2}\in L^{1}(\overline{\Gamma}_{N}), a3L1(Γ^N)a_{3}\in L^{1}(\widehat{\Gamma}_{N}), c1L(Γ¯N)c_{1}\in L^{\infty}(\overline{\Gamma}_{N}) and c2L(Γ^N)c_{2}\in L^{\infty}(\widehat{\Gamma}_{N}). In our current setting we have hΩ=h¯=0h_{\Omega}=\overline{h}=0 and h^(𝐬,𝐮^)=12𝒳Γtar(𝐬)W(𝐮^𝐮tar)(𝐮^𝐮tar)\widehat{h}(\bm{s},{\widehat{{\bf u}}})=\frac{1}{2}\mathrm{\mathcal{X}}_{\Gamma^{\rm tar}}(\bm{s})W({\widehat{{\bf u}}}-\bm{u}^{\rm tar})\cdot({\widehat{{\bf u}}}-\bm{u}^{\rm tar}).

Remark 2.2.

It is also possible to consider mass constraints for 𝛗{\bm{\varphi}} of the form

1|Ω|Ω𝝋dx𝜶 or 1|Ω|Ω𝝋dx𝜷 or 𝜷1|Ω|Ω𝝋dx𝜶,\displaystyle\frac{1}{|\Omega|}\int_{\Omega}{\bm{\varphi}}\,\mathrm{dx}\leq\bm{\alpha}\quad\text{ or }\quad\frac{1}{|\Omega|}\int_{\Omega}{\bm{\varphi}}\,\mathrm{dx}\geq\bm{\beta}\quad\text{ or }\quad\bm{\beta}\leq\frac{1}{|\Omega|}\int_{\Omega}{\bm{\varphi}}\,\mathrm{dx}\leq\bm{\alpha},

for fixed vectors 𝛂,𝛃ΔL\bm{\alpha},\bm{\beta}\in\Delta^{L} (possibly also 𝛂=𝛃\bm{\alpha}=\bm{\beta}), where in the above the inequalities are taken component-wise. These are convex constraints and thus when included into the definition of 𝒰ad\mathcal{U}_{\rm ad}, the design space remains a closed and convex set. Then, in the corresponding necessary optimality condition, associated Lagrange multipliers will appear, see [13, 14] for more details.

Notation.

For a Banach space XX, we denote its topological dual by XX^{*}, and the corresponding duality pairing by ,X\langle\cdot,\cdot\rangle_{X}. For any p[1,]p\in[1,\infty] and k>0k>0, the standard Lebesgue and Sobolev spaces over Ω\Omega are denoted by Lp:=Lp(Ω)L^{p}:=L^{p}(\Omega) and Wk,p:=Wk,p(Ω)W^{k,p}:=W^{k,p}(\Omega) with the corresponding norms Lp(Ω)\|\cdot\|_{L^{p}(\Omega)} and Wk,p(Ω)\|\cdot\|_{W^{k,p}(\Omega)}. In the special case p=2p=2, these become Hilbert spaces and we employ the notation Hk:=Hk(Ω)=Wk,2(Ω)H^{k}:=H^{k}(\Omega)=W^{k,2}(\Omega) with the corresponding norm Hk(Ω)\|\cdot\|_{H^{k}(\Omega)}. For our subsequent analysis, we introduce the spaces

H¯D1(Ω,d):={𝐯H1(Ω,d):𝐯=𝟎a.e. on Γ¯D},\displaystyle\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}):=\left\{{\bf v}\in H^{1}(\Omega,{\mathbb{R}}^{d}):{\bf v}=\bm{0}\quad\text{a.e. on $\overline{\Gamma}_{D}$}\right\},

and

H^D1(Ω,d):={𝐯H1(Ω,d):𝐯=𝟎a.e. on Γ^D}.\displaystyle\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}):=\left\{{\bf v}\in H^{1}(\Omega,{\mathbb{R}}^{d}):{\bf v}=\bm{0}\quad\text{a.e. on $\widehat{\Gamma}_{D}$}\right\}.

For brevity, the corresponding norms are denoted by the same symbol H1(Ω)\mathopen{\|}\cdot\mathclose{\|}_{H^{1}(\Omega)} if no confusion may arise. Vectors, matrices, and vector- or matrix-valued functions will be denoted by bold symbols. Furthermore, for a subset ΓNΓ\Gamma_{N}\subset\Gamma, we consider the function space

H001/2(ΓN,d):={𝐯H1/2(ΓN,d):𝐯~H1/2(Γ,d)},H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d}):=\left\{{\bf v}\in H^{1/2}(\Gamma_{N},{\mathbb{R}}^{d}):\widetilde{\bf v}\in H^{1/2}(\Gamma,{\mathbb{R}}^{d})\right\},

where 𝐯~\widetilde{\bf v} denotes the trivial extension of 𝐯{\bf v} to Γ\Gamma, and we endow it with the norm

𝐯H001/2(ΓN):=𝐯~H1/2(Γ),𝐯H001/2(ΓN,d).\displaystyle\mathopen{\|}{\bf v}\mathclose{\|}_{H^{1/2}_{00}(\Gamma_{N})}:=\mathopen{\|}\widetilde{\bf v}\mathclose{\|}_{H^{1/2}(\Gamma)},\quad{\bf v}\in H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d}).

We highlight that the above definition is not redundant as in general the trivial extension of a H1/2(ΓN,d)H^{1/2}(\Gamma_{N},{\mathbb{R}}^{d}) function does not belong to H1/2(Γ,d)H^{1/2}(\Gamma,{\mathbb{R}}^{d}). Besides, we remark that H001/2(ΓN,d)H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d}) is a Hilbert space and

(H001/2(ΓN,d),L2(ΓN,d),H001/2(ΓN,d))\displaystyle(H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d}),L^{2}(\Gamma_{N},{\mathbb{R}}^{d}),H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d})^{*})

forms a Hilbert triple (see, e.g., [48]).

For the forthcoming analysis we make the following structural assumptions.

  1. (A1)

    The domain Ωd\Omega\subset{\mathbb{R}}^{d}, d{2,3}d\in\{2,3\}, is a bounded domain with C1,1C^{1,1} or convex boundary Γ=Ω\Gamma=\partial\Omega that admits a partition Γ=cl(Γ¯D)cl(Γ¯N)\Gamma={\rm cl}(\overline{\Gamma}_{D})\cup{\rm cl}(\overline{\Gamma}_{N}) with relative open subsets Γ¯D\overline{\Gamma}_{D} and Γ¯N\overline{\Gamma}_{N} such that Γ¯DΓ¯N=\overline{\Gamma}_{D}\cap\overline{\Gamma}_{N}=\emptyset and Γ¯D\overline{\Gamma}_{D}\neq\emptyset. Here, cl(A){\rm cl}(A) denotes the closure of the set AA. The same assumptions are made for Γ^D\widehat{\Gamma}_{D} and Γ^N\widehat{\Gamma}_{N}.

  2. (A2)

    The elasticity tensor ¯{\overline{{\mathbb{C}}}} is assumed to be a tensor-valued function

    ¯:Ld×d×d×d,\displaystyle{\overline{{\mathbb{C}}}}:{\mathbb{R}}^{L}\to{\mathbb{R}}^{d\times d\times d\times d},

    with ¯ijklC1,1(L,){\overline{{\mathbb{C}}}}_{ijkl}\in{C^{1,1}({\mathbb{R}}^{L},{\mathbb{R}})}, i,j,k,l{1,,d}i,j,k,l\in\{1,\dots,d\}. Moreover, it fulfils the symmetry conditions

    ¯ijkl=¯klij=¯ijlk=¯jiklfor all i,j,k,l{1,,d},\displaystyle{\overline{{\mathbb{C}}}}_{ijkl}={\overline{{\mathbb{C}}}}_{klij}={\overline{{\mathbb{C}}}}_{ijlk}={\overline{{\mathbb{C}}}}_{jikl}\quad\text{for all $i,j,k,l\in\{1,\dots,d\}$,}

    and there exist positive constants C0,C1C_{0},C_{1}, and C2C_{2} such that, for all 𝝋,𝐡L{\bm{\varphi}},{\bf h}\in{\mathbb{R}}^{L},

    C0|𝐀|2¯(𝝋)𝐀:𝐀C1|𝐀|2\displaystyle C_{0}|{{\bf A}}|^{2}\leq{\overline{{\mathbb{C}}}}({\bm{\varphi}}){{\bf A}}:{{\bf A}}\leq C_{1}|{{\bf A}}|^{2}\qquad 𝐀symd×d{𝟎},\displaystyle\forall{{\bf A}}\in{\mathbb{R}}^{d\times d}_{\rm sym}\setminus\{\bm{0}\}, (2.4)
    |¯(𝝋)𝐡𝐀:𝐁|C2|𝐡||𝐀||𝐁|\displaystyle|{\overline{{\mathbb{C}}}}^{\prime}({\bm{\varphi}}){\bf h}\,{\bf A}:{\bf B}|\leq C_{2}|{\bf h}||{\bf A}||{\bf B}|\qquad 𝐀,𝐁symd×d{𝟎},\displaystyle\forall{{\bf A},{\bf B}}\in{\mathbb{R}}^{d\times d}_{\rm sym}\setminus\{\bm{0}\}, (2.5)

    where 𝐀:𝐁=i,j=1nAijBij{\bf A}:{\bf B}=\sum_{i,j=1}^{n}A_{ij}B_{ij}, |𝐀|=𝐀:𝐀|{\bf A}|=\sqrt{{\bf A}:{\bf A}}, and for every 𝐡=(h1,,hL)L{\bf h}=(h_{1},\dots,h_{L})\in{\mathbb{R}}^{L},

    [¯(𝝋)𝐡]ijkl:=m=1Lm¯ijkl(𝝋)hm for all i,j,k,l{1,,d}.\displaystyle[{\overline{{\mathbb{C}}}}^{\prime}({\bm{\varphi}}){\bf h}]_{ijkl}:=\sum_{m=1}^{L}\partial_{m}{\overline{{\mathbb{C}}}}_{ijkl}({\bm{\varphi}})h_{m}\quad\text{ for all }i,j,k,l\in\{1,\dots,d\}.

    The set symd×d{\mathbb{R}}^{d\times d}_{\rm sym} consists of the symmetric (d×d)({d\times d})-matrices. The same assumptions are made for ^\widehat{{\mathbb{C}}}.

  3. (A3)

    The multiwell potential Ψ\Psi possesses the form

    Ψ:L{+},Ψ=Ψ~+IΔ,\displaystyle\Psi:{\mathbb{R}}^{L}\to{\mathbb{R}}\cup\{+\infty\},\quad\Psi=\widetilde{\Psi}+I_{\Delta},

    where Ψ~C1,1(L)\widetilde{\Psi}\in{C^{1,1}({\mathbb{R}}^{L})} and the indicator function IΔI_{\Delta} of the simplex ΔL\Delta^{L} is defined as

    IΔ(𝝋)={0if 𝝋ΔL,+otherwise.\displaystyle I_{\Delta}({\bm{\varphi}})=\begin{cases}0\quad&\text{if ${\bm{\varphi}}\in\Delta^{L}$},\\ +\infty\quad&\text{otherwise}.\end{cases}
  4. (A4)

    The function χ:L\chi:{\mathbb{R}}^{L}\to{\mathbb{R}} is C1,1(L){C^{1,1}({\mathbb{R}}^{L})} and there exist a positive constant χ0\chi_{0} such that

    0χ(𝝋)χ0 for all 𝝋L.\displaystyle 0\leq\chi({\bm{\varphi}})\leq\chi_{0}\qquad\text{ for all }{\bm{\varphi}}\in{\mathbb{R}}^{L}.
  5. (A5)

    The data of the problems satisfy

    𝐅¯,𝐅^\displaystyle\overline{{\bf F}},\widehat{{\bf F}} L2(Ω,d),𝐔¯H1/2(Γ¯D,d),𝐔^H1/2(Γ^D,d),\displaystyle\in L^{2}(\Omega,{\mathbb{R}}^{d}),\quad\overline{{\bf U}}\in H^{1/2}(\overline{\Gamma}_{D},{\mathbb{R}}^{d}),\quad\widehat{{\bf U}}\in H^{1/2}(\widehat{\Gamma}_{D},{\mathbb{R}}^{d}),
    𝐠¯\displaystyle\overline{{\bf g}} H001/2(Γ¯N,d),𝐠^H001/2(Γ^N,d).\displaystyle\in H^{1/2}_{00}(\overline{\Gamma}_{N},{\mathbb{R}}^{d})^{*},\quad\widehat{{\bf g}}\in H^{1/2}_{00}(\widehat{\Gamma}_{N},{\mathbb{R}}^{d})^{*}.
  6. (A6)

    The target displacement 𝐮tarL2(Γtar,d){\bf u}^{\rm tar}\in L^{2}(\Gamma^{\rm tar},{\mathbb{R}}^{d}), where ΓtarΓ^N\Gamma^{\rm tar}\subset\widehat{\Gamma}_{N}.

It is worth pointing out that condition (A3) entails that Ψ(𝝋)=Ψ~(𝝋)\Psi({\bm{\varphi}})=\widetilde{\Psi}({\bm{\varphi}}) for every 𝝋𝒰ad{\bm{\varphi}}\in\mathcal{U}_{\rm ad}.

3 Analysis of the design optimisation problem

3.1 Linear elasticity system with mixed boundary conditions

In this section we provide a preliminary well-posedness result for the following linear elasticity system with mixed boundary conditions

div((𝐮)+𝔽)\displaystyle-\mathop{\rm div}\nolimits\big{(}{\mathbb{C}}\mathcal{E}({\bf u})+{\mathbb{F}}\big{)} =𝐟\displaystyle={\bf f}\qquad in Ω,\displaystyle\text{in $\Omega$}, (3.1a)
𝐮\displaystyle{\bf u} =𝟎\displaystyle=\bm{0}\qquad on ΓD,\displaystyle\text{on $\Gamma_{D}$}, (3.1b)
((𝐮)+𝔽)𝐧\displaystyle({\mathbb{C}}\mathcal{E}({\bf u})+{\mathbb{F}}){\bf n} =𝐠\displaystyle={\bf g}\qquad on ΓN.\displaystyle\text{on $\Gamma_{N}$}. (3.1c)

The well-posedness of the system (3.1) is formulated as follows.

Proposition 3.1.

In addition to (A1), suppose that L(Ω,d×d×d×d){\mathbb{C}}\in L^{\infty}(\Omega,{\mathbb{R}}^{d\times d\times d\times d}) fulfils the symmetry conditions

ijkl\displaystyle{\mathbb{C}}_{ijkl} =klij=ijlk=jikl\displaystyle={\mathbb{C}}_{klij}={\mathbb{C}}_{ijlk}={\mathbb{C}}_{jikl}\quad for all i,j,k,l{1,,d}i,j,k,l\in\{1,\dots,d\},

and satisfies

λ|𝐀|2(𝐀):𝐀Λ|𝐀|2 for all 𝐀symd×d{𝟎}and a.e. in Ω,\displaystyle\lambda|{{\bf A}}|^{2}\leq({\mathbb{C}}{{\bf A}}):{{\bf A}}\leq\Lambda|{{\bf A}}|^{2}\quad\text{ for all }{\bf A}\in{\mathbb{R}}^{d\times d}_{\rm sym}\setminus\{\bm{0}\}\,\,\text{and a.e. in $\Omega$}, (3.2)

for some positive constants λ\lambda and Λ\Lambda. Then, for every 𝔽L2(Ω,d×d){\mathbb{F}}\in L^{2}(\Omega,{\mathbb{R}}^{d\times d}), 𝐟L2(Ω,d){\bf f}\in L^{2}(\Omega,{\mathbb{R}}^{d}) and 𝐠H001/2(ΓN,d){\bf g}\in H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d})^{*}, there exists a unique weak solution 𝐮HD1(Ω,d):={𝐯H1(Ω,d):𝐯=𝟎 a.e. in ΓD}{\bf u}\in H^{1}_{D}(\Omega,{\mathbb{R}}^{d}):=\{\bm{v}\in H^{1}(\Omega,{\mathbb{R}}^{d})\,:\,\bm{v}=\bm{0}\text{ a.e.~{}in }\Gamma_{D}\} to the elasticity system (3.1) in the sense that

Ω(𝐮):(𝐯)dx=Ω(𝔽:(𝐯)𝐟𝐯)dx+𝐠,𝐯H001/2(ΓN)𝐯HD1(Ω,d),\displaystyle\int_{\Omega}{\mathbb{C}}\mathcal{E}({\bf u}):\mathcal{E}({\bf v})\,\mathrm{dx}=-\int_{\Omega}({\mathbb{F}}:\mathcal{E}({\bf v})-{\bf f}\cdot{\bf v})\,\mathrm{dx}+\mathopen{\langle}{\bf g},{\bf v}\mathclose{\rangle}_{H^{1/2}_{00}(\Gamma_{N})}\quad\forall{\bf v}\in H^{1}_{D}(\Omega,{\mathbb{R}}^{d}), (3.3)

and there exists a positive constant CC independent of 𝐮{\bf u} such that

𝐮H1(Ω)C(𝔽L2(Ω)+𝐟L2(Ω)+𝐠H001/2(ΓN)).\displaystyle\mathopen{\|}{\bf u}\mathclose{\|}_{H^{1}(\Omega)}\leq C\left(\mathopen{\|}{\mathbb{F}}\mathclose{\|}_{L^{2}(\Omega)}+\mathopen{\|}{\bf f}\mathclose{\|}_{L^{2}(\Omega)}+\mathopen{\|}{\bf g}\mathclose{\|}_{H^{1/2}_{00}(\Gamma_{N})^{*}}\right). (3.4)

Note that whenever 𝐠L2(ΓN,d)H001/2(ΓN,d){\bf g}\in L^{2}(\Gamma_{N},{\mathbb{R}}^{d})\subset H^{1/2}_{00}(\Gamma_{N},{\mathbb{R}}^{d})^{*}, we can identify the duality product in (3.3) as the standard boundary integral, that is,

𝐠,𝐯H001/2(ΓN)=ΓN𝐠𝐯dd1.\displaystyle\mathopen{\langle}{\bf g},{\bf v}\mathclose{\rangle}_{H^{1/2}_{00}(\Gamma_{N})}=\int_{\Gamma_{N}}{\bf g}\cdot{\bf v}\,\mathrm{d}\mathcal{H}^{d-1}.
Proof of Proposition 3.1.

The variational equality (3.3) admits a unique solution by a direct application of the Lax–Milgram theorem (cf., e.g., [5]). In this direction, we set

𝐕=HD1(Ω,d)\displaystyle{\bf V}=H^{1}_{D}(\Omega,{\mathbb{R}}^{d})
a(,):𝐕×𝐕,a(𝐮,𝐯):=Ω(𝐮):(𝐯)dx,\displaystyle a(\cdot,\cdot):{\bf V}\times{\bf V}\to{\mathbb{R}},\quad a({\bf u},{\bf v}):=\int_{\Omega}{\mathbb{C}}\mathcal{E}({\bf u}):\mathcal{E}({\bf v})\,\mathrm{dx},
𝐅,𝐯:=Ω(𝔽:(𝐯)𝐟𝐯)dx+𝐠,𝐯H001/2(ΓN).\displaystyle\langle{\bf F},{\bf v}\rangle:=-\int_{\Omega}({\mathbb{F}}:\mathcal{E}({\bf v})-{\bf f}\cdot{\bf v})\,\mathrm{dx}+\mathopen{\langle}{\bf g},{\bf v}\mathclose{\rangle}_{H^{1/2}_{00}(\Gamma_{N})}.

It is worth noticing that it readily follows from the assumptions on 𝔽{\mathbb{F}}, 𝐟{\bf f}, and 𝐠{\bf g} that 𝐅𝐕{\bf F}\in{\bf V}^{*}. With the above notation, (3.3) can then be rewritten as the variational problem

a(𝐮,𝐯)=𝐅,𝐯𝐯𝐕.\displaystyle a({\bf u},{\bf v})=\mathopen{\langle}{\bf F},{\bf v}\mathclose{\rangle}\quad\forall{\bf v}\in{\bf V}.

Thus, to apply the Lax–Milgram theorem, it is sufficient to show the bilinear form a(,)a(\cdot,\cdot) is continuous and coercive in 𝐕{\bf V}. By (3.2) we have

|a(𝐮,𝐯)|C𝐮H1(Ω)𝐯H1(Ω)C𝐮𝐕𝐯𝐕𝐮,𝐯𝐕,\displaystyle|a({\bf u},{\bf v})|\leq C\mathopen{\|}{\bf u}\mathclose{\|}_{H^{1}(\Omega)}\mathopen{\|}{\bf v}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}{\bf u}\mathclose{\|}_{{\bf V}}\mathopen{\|}{\bf v}\mathclose{\|}_{{\bf V}}\quad\forall{\bf u},{\bf v}\in{{\bf V}},

while (3.2) and Korn’s inequality yield the 𝐕{\bf V}-coercivity:

|a(𝐯,𝐯)|λ(𝐯)L2(Ω)2C(λ,CK)𝐯𝐕2𝐯𝐕,\displaystyle|a({\bf v},{\bf v})|\geq\lambda\mathopen{\|}\mathcal{E}({\bf v})\mathclose{\|}_{L^{2}(\Omega)}^{2}\geq C(\lambda,{C_{K}})\mathopen{\|}{\bf v}\mathclose{\|}_{{\bf V}}^{2}\quad\forall{\bf v}\in{{\bf V}},

with a constant CK{C_{K}} arising from Korn’s inequality. Thus, the existence and uniqueness of 𝐮HD1(Ω,d){\bf u}\in H^{1}_{D}(\Omega,{\mathbb{R}}^{d}) solving (3.3), as well as the estimate (3.4), readily follow from the Lax–Milgram theorem. ∎

We end this section with another abstract result that will be useful for the subsequent analysis. Consider the following problem with inhomogeneous data on the Dirichlet boundary:

div((𝐮))\displaystyle-\mathop{\rm div}\nolimits\big{(}{\mathbb{C}}\mathcal{E}({\bf u})\big{)} =𝟎\displaystyle=\bm{0}\qquad in Ω,\displaystyle\text{ in }\Omega, (3.5)
𝐮\displaystyle{\bf u} =𝒰\displaystyle={\cal U}\qquad on ΓD,\displaystyle\text{ on }\Gamma_{D},
((𝐮))𝐧\displaystyle\big{(}{\mathbb{C}}\mathcal{E}({\bf u})\big{)}{\bf n} =𝟎\displaystyle=\bm{0}\qquad on ΓN.\displaystyle\text{ on }\Gamma_{N}.

Well-known theory yields that, for every 𝒰H1/2(ΓD,d){\cal U}\in H^{1/2}(\Gamma_{D},{\mathbb{R}}^{d}), there exists a unique weak solution 𝐮H1(Ω,d){\bf u}\in H^{1}(\Omega,{\mathbb{R}}^{d}). The proof follows similarly to the above as a direct consequence of the Lax–Milgram theorem. This allows us to introduce the associated solution operator, that we call the extension operator

:H1/2(ΓD,d)H1(Ω,d),:𝒰𝐮,\displaystyle{\cal H}:H^{1/2}(\Gamma_{D},{\mathbb{R}}^{d})\to H^{1}(\Omega,{\mathbb{R}}^{d}),\quad{\cal H}:{\cal U}\mapsto{\bf u},\quad (3.6)

where 𝐮{\bf u} is the unique weak solution to system (3.5).

3.2 Well-posedness of the state systems

Similarly to (3.5) and (3.6), we can introduce the extension operators ¯\overline{\cal H} and ^\widehat{\cal H} related to ¯\overline{{\mathbb{C}}}, Γ¯D\overline{\Gamma}_{D} and ^\widehat{{\mathbb{C}}}, Γ^D\widehat{\Gamma}_{D}, respectively. Then defining the functions H¯:=¯(𝑼¯),H^:=^(𝑼^)H1(Ω,d)\overline{H}:=\overline{\cal H}(\overline{\bm{U}}),\widehat{H}:=\widehat{\cal H}(\widehat{\bm{U}})\in H^{1}(\Omega,{\mathbb{R}}^{d}) allows us to transform (2.1) and (2.2) into the equivalent problems

div(¯(𝝋)(𝐮¯new+H¯))\displaystyle-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}^{\rm new}+{\overline{H}})\big{)} =𝐅¯\displaystyle=\overline{{\bf F}}\qquad in Ω,\displaystyle\text{ in }\Omega, (3.7)
𝐮¯new\displaystyle{\overline{{\bf u}}}^{\rm new} =𝟎\displaystyle=\bm{0}\qquad on Γ¯D,\displaystyle\text{ on }\overline{\Gamma}_{D},
(¯(𝝋)(𝐮¯new+H¯))𝐧\displaystyle\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}^{\rm new}+{\overline{H}})\big{)}{\bf n} =𝐠¯\displaystyle=\overline{{\bf g}}\qquad on Γ¯N,\displaystyle\text{ on }\overline{\Gamma}_{N},

and

div(^(𝝋)((𝐮^new+H^)χ(𝝋)(𝐮¯new+H¯))\displaystyle-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf u}}}^{\rm new}+{\widehat{H}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}^{\rm new}+{\overline{H}})\big{)} =𝐅^\displaystyle=\widehat{{\bf F}}\qquad in Ω,\displaystyle\text{ in }\Omega, (3.8)
𝐮^new\displaystyle{\widehat{{\bf u}}}^{\rm new} =𝟎\displaystyle=\bm{0}\qquad on Γ^D,\displaystyle\text{ on }\widehat{\Gamma}_{D},
(^(𝝋)((𝐮^new+H^)χ(𝝋)(𝐮¯new+H¯))𝐧\displaystyle\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf u}}}^{\rm new}+{\widehat{H}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}^{\rm new}+{\overline{H}})\big{)}{\bf n} =𝐠^\displaystyle=\widehat{{\bf g}}\qquad on Γ^N,\displaystyle\text{ on }\widehat{\Gamma}_{N},

where we set

𝐮¯=𝐮¯new+H¯,𝐮^=𝐮^new+H^.{\overline{{\bf u}}}={\overline{{\bf u}}}^{\rm new}+{\overline{H}},\quad{\widehat{{\bf u}}}={\widehat{{\bf u}}}^{\rm new}+{\widehat{H}}.

For a cleaner presentation, we abuse notation and use the same variables 𝐮¯{\overline{{\bf u}}} and 𝐮^{\widehat{{\bf u}}} to denote 𝐮¯new{\overline{{\bf u}}}^{\rm new} and 𝐮^new{\widehat{{\bf u}}}^{\rm new}. Then, the well-posedness of (2.1) and (2.2) (equivalently (3.7) and (3.8)) are formulated as follows.

Theorem 3.1.

Under (A1)(A5), for every 𝛗L(Ω,L){\bm{\varphi}}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}), there exists a unique solution pair (𝐮¯,𝐮^)H¯D1(Ω,d)×H^D1(Ω,d)({\overline{{\bf u}}},{\widehat{{\bf u}}})\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) satisfying

Ω¯(𝝋)(𝐮¯):(𝜻¯)dx=Ω𝐅¯𝜻¯dxΩ¯(𝝋)(H¯):(𝜻¯)dx+𝐠¯,𝜻¯H001/2(Γ¯N),\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}(\overline{{\bm{\zeta}}})\,\mathrm{dx}=\int_{\Omega}\overline{{\bf F}}\cdot\overline{{\bm{\zeta}}}\,\mathrm{dx}-\int_{\Omega}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{H}}):\mathcal{E}(\overline{{\bm{\zeta}}})\,\mathrm{dx}+\mathopen{\langle}\overline{{\bf g}},\overline{{\bm{\zeta}}}\mathclose{\rangle}_{H^{1/2}_{00}(\overline{\Gamma}_{N})}, (3.9)
Ω^(𝝋)(𝐮^):(𝜻^)dxΩ^(𝝋)χ(𝝋)(𝐮¯):(𝜻^)dx=Ω𝐅^𝜻^dx\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\widehat{{\bf u}}}):\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}=\int_{\Omega}\widehat{{\bf F}}\cdot\widehat{{\bm{\zeta}}}\,\mathrm{dx}
+Ω^(𝝋)(H^):(𝜻^)dxΩ^(𝝋)χ(𝝋)(H¯):(𝜻^)dx+𝐠^,𝜻^H001/2(Γ^N)\displaystyle\quad+\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\widehat{H}}):\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}({\overline{H}}):\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}+\mathopen{\langle}\widehat{{\bf g}},\widehat{{\bm{\zeta}}}\mathclose{\rangle}_{H^{1/2}_{00}(\widehat{\Gamma}_{N})} (3.10)

for all 𝛇¯H¯D1(Ω,d)\overline{{\bm{\zeta}}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝛇^H^D1(Ω,d)\widehat{{\bm{\zeta}}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Moreover, there exists a positive constant CC, independent of 𝛗{\bm{\varphi}}, such that

𝐮¯H1(Ω)+𝐮^H1(Ω)C.\displaystyle\mathopen{\|}{\overline{{\bf u}}}\mathclose{\|}_{H^{1}(\Omega)}+\mathopen{\|}{\widehat{{\bf u}}}\mathclose{\|}_{H^{1}(\Omega)}\leq C. (3.11)
Proof.

For (3.9) we invoke Proposition 3.1 with the specifications

𝐮=𝐮¯,=¯(𝝋),𝔽=¯(𝝋)(H¯),𝐟=𝐅¯,𝐠=𝐠¯,ΓD=Γ¯D,ΓN=Γ¯N,\displaystyle{\bf u}={\overline{{\bf u}}},\quad{\mathbb{C}}=\overline{{\mathbb{C}}}({\bm{\varphi}}),\quad{\mathbb{F}}=\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{H}}),\quad{\bf f}=\overline{{\bf F}},\quad{\bf g}=\overline{{\bf g}},\quad\Gamma_{D}=\overline{\Gamma}_{D},\quad\Gamma_{N}=\overline{\Gamma}_{N},

and for (3.10) we consider

𝐮=𝐮^,=^(𝝋),𝔽=^(𝝋)((H^)χ(𝝋)(𝐮¯+H¯)),𝐟=𝐅^,𝐠=𝐠^,\displaystyle{\bf u}={\widehat{{\bf u}}},\quad{\mathbb{C}}=\widehat{{\mathbb{C}}}({\bm{\varphi}}),\quad{\mathbb{F}}=\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{H}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}+{\overline{H}})),\quad{\bf f}=\widehat{{\bf F}},\quad{\bf g}=\widehat{{\bf g}},
ΓD=Γ^D,ΓN=Γ^N,\displaystyle\Gamma_{D}=\widehat{\Gamma}_{D},\quad\Gamma_{N}=\widehat{\Gamma}_{N},

to obtain the existence and uniqueness of solutions 𝐮¯{\overline{{\bf u}}} and 𝐮^{\widehat{{\bf u}}}. Lastly the estimate (3.11) can be obtained from (3.4) and the uniform boundedness of the tensors ¯\overline{{\mathbb{C}}} and ^\widehat{{\mathbb{C}}} in (A2). ∎

Theorem 3.2.

Under (A1)(A5), for i=1,2i=1,2, let 𝛗iL(Ω,L){\bm{\varphi}}_{i}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}) with 𝛗iL(Ω)R\mathopen{\|}{\bm{\varphi}}_{i}\mathclose{\|}_{L^{\infty}(\Omega)}\leq R, where R>0R>0 is fixed, and let (𝐮¯i,𝐮^i)({\overline{{\bf u}}}_{i},{\widehat{{\bf u}}}_{i}) denote the unique solutions to systems (3.7) and (3.8) corresponding to 𝛗i{\bm{\varphi}}_{i}, but with the same data 𝐅¯{\overline{{\bf F}}}, 𝐅^\widehat{{\bf F}}, 𝐠¯\overline{{\bf g}}, 𝐠^\widehat{{\bf g}}, H¯{\overline{H}} and H^{\widehat{H}}. Then, there exists a positive constant CC, independent of the differences, such that

𝐮¯1𝐮¯2H1(Ω)\displaystyle\mathopen{\|}{\overline{{\bf u}}}_{1}-{\overline{{\bf u}}}_{2}\mathclose{\|}_{H^{1}(\Omega)} C𝝋1𝝋2L(Ω),\displaystyle\leq C\mathopen{\|}{\bm{\varphi}}_{1}-{\bm{\varphi}}_{2}\mathclose{\|}_{L^{\infty}(\Omega)}, (3.12)
𝐮^1𝐮^2H1(Ω)\displaystyle\mathopen{\|}{\widehat{{\bf u}}}_{1}-{\widehat{{\bf u}}}_{2}\mathclose{\|}_{H^{1}(\Omega)} C(𝐮¯1𝐮¯2)L2(Ω)+C𝝋1𝝋2L(Ω).\displaystyle\leq C\mathopen{\|}\mathcal{E}({\overline{{\bf u}}}_{1}-{\overline{{\bf u}}}_{2})\mathclose{\|}_{L^{2}(\Omega)}+C\mathopen{\|}{\bm{\varphi}}_{1}-{\bm{\varphi}}_{2}\mathclose{\|}_{L^{\infty}(\Omega)}.
Proof.

To start, let us set

𝝋:=𝝋1𝝋2,𝐮¯:=𝐮¯1𝐮¯2,𝐮^:=𝐮^1𝐮^2.\displaystyle{\bm{\varphi}}:={\bm{\varphi}}_{1}-{\bm{\varphi}}_{2},\quad{\overline{{\bf u}}}:={\overline{{\bf u}}}_{1}-{\overline{{\bf u}}}_{2},\quad{\widehat{{\bf u}}}:={\widehat{{\bf u}}}_{1}-{\widehat{{\bf u}}}_{2}.

Then, we consider the difference between the variational equalities (3.9)–(3.10) written for (𝝋1,𝐮¯1,𝐮^1)({\bm{\varphi}}_{1},{\overline{{\bf u}}}_{1},{\widehat{{\bf u}}}_{1}) and for (𝝋2,𝐮¯2,𝐮^2)({\bm{\varphi}}_{2},{\overline{{\bf u}}}_{2},{\widehat{{\bf u}}}_{2}) to infer that

Ω(¯(𝝋1)(𝐮¯1)¯(𝝋2)(𝐮¯2)):(𝜻¯)dx=0,\displaystyle\int_{\Omega}\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}}_{1})\mathcal{E}({\overline{{\bf u}}}_{1})-\overline{{\mathbb{C}}}({\bm{\varphi}}_{2})\mathcal{E}({\overline{{\bf u}}}_{2})\big{)}:\mathcal{E}(\overline{{\bm{\zeta}}})\,\mathrm{dx}=0, (3.13)
Ω(^(𝝋1)(𝐮^1)^(𝝋2)(𝐮^2)):(𝜻^)dx\displaystyle\int_{\Omega}\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}}_{1})\mathcal{E}({\widehat{{\bf u}}}_{1})-\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2})\mathcal{E}({\widehat{{\bf u}}}_{2})\big{)}:\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}
Ω(^(𝝋1)χ(𝝋1)(𝐮¯1)^(𝝋2)χ(𝝋2)(𝐮¯2)):(𝜻^)dx=0,\displaystyle\quad-\int_{\Omega}\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}}_{1})\chi({\bm{\varphi}}_{1})\mathcal{E}({\overline{{\bf u}}}_{1})-\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2})\chi({\bm{\varphi}}_{2})\mathcal{E}({\overline{{\bf u}}}_{2})\big{)}:\mathcal{E}(\widehat{{\bm{\zeta}}})\,\mathrm{dx}=0, (3.14)

for all 𝜻¯H¯D1(Ω,d)\overline{{\bm{\zeta}}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝜻^H^D1(Ω,d).\widehat{{\bm{\zeta}}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Choosing 𝜻¯=𝐮¯\overline{{\bm{\zeta}}}={\overline{{\bf u}}} and invoking condition (2.4) yields

C0(𝐮¯)L2(Ω)2\displaystyle C_{0}\mathopen{\|}\mathcal{E}({\overline{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}^{2} Ω(¯(𝝋1)¯(𝝋2))(𝐮¯2):(𝐮¯)dx\displaystyle\leq-\int_{\Omega}(\overline{{\mathbb{C}}}({\bm{\varphi}}_{1})-\overline{{\mathbb{C}}}({\bm{\varphi}}_{2}))\mathcal{E}({\overline{{\bf u}}}_{2}):\mathcal{E}({\overline{{\bf u}}})\,\mathrm{dx} (3.15)
C02(𝐮¯)L2(Ω)2+C𝝋L(Ω)2.\displaystyle\leq\frac{C_{0}}{2}\mathopen{\|}\mathcal{E}({\overline{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}^{2}+C\mathopen{\|}{\bm{\varphi}}\mathclose{\|}_{L^{\infty}(\Omega)}^{2}.

By Korn’s inequality we infer

𝐮¯1𝐮¯2H1(Ω)C𝝋1𝝋2L(Ω).\displaystyle\mathopen{\|}{\overline{{\bf u}}}_{1}-{\overline{{\bf u}}}_{2}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}{\bm{\varphi}}_{1}-{\bm{\varphi}}_{2}\mathclose{\|}_{L^{\infty}(\Omega)}. (3.16)

Then, inserting 𝜻^=𝐮^\widehat{{\bm{\zeta}}}={\widehat{{\bf u}}} in (3.14), and invoking the Lipschitz continuity of ^\widehat{{\mathbb{C}}} and χ\chi from (A2) and (A4), as well as (3.16), yields

C0(𝐮^)L2(Ω)2\displaystyle C_{0}\mathopen{\|}\mathcal{E}({\widehat{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}^{2} (3.17)
Ω(^(𝝋1)^(𝝋2))(𝐮^2):(𝐮^)+(^(𝝋1)^(𝝋2))χ(𝝋1)(𝐮¯1):(𝐮^)dx\displaystyle\quad\leq-\int_{\Omega}(\widehat{{\mathbb{C}}}({\bm{\varphi}}_{1})-\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2}))\mathcal{E}({\widehat{{\bf u}}}_{2}):\mathcal{E}({\widehat{{\bf u}}})+(\widehat{{\mathbb{C}}}({\bm{\varphi}}_{1})-\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2}))\chi({\bm{\varphi}}_{1})\mathcal{E}({\overline{{\bf u}}}_{1}):\mathcal{E}({\widehat{{\bf u}}})\,\mathrm{dx}
Ω^(𝝋2)(χ(𝝋1)χ(𝝋2))(𝐮¯1):(𝐮^)dxΩ^(𝝋2)χ(𝝋2)(𝐮¯):(𝐮^)dx\displaystyle\qquad-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2})(\chi({\bm{\varphi}}_{1})-\chi({\bm{\varphi}}_{2}))\mathcal{E}({\overline{{\bf u}}}_{1}):\mathcal{E}({\widehat{{\bf u}}})\,\mathrm{dx}-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}}_{2})\chi({\bm{\varphi}}_{2})\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\widehat{{\bf u}}})\,\mathrm{dx}
C02(𝐮^)L2(Ω)2+C((𝐮¯)L2(Ω)2+𝝋L(Ω)2).\displaystyle\quad\leq\frac{C_{0}}{2}\mathopen{\|}\mathcal{E}({\widehat{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}^{2}+C\Big{(}\mathopen{\|}\mathcal{E}({\overline{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}^{2}+\mathopen{\|}{\bm{\varphi}}\mathclose{\|}_{L^{\infty}(\Omega)}^{2}\Big{)}.

Applying Korn’s inequality leads to (3.12). ∎

Corollary 3.1.

Suppose that (A1)(A5) hold. Let R>0R>0 and let {𝛗n}n\{{\bm{\varphi}}_{n}\}_{n\in{\mathbb{N}}} be a sequence of functions in L(Ω,L)L^{\infty}(\Omega,{\mathbb{R}}^{L}) such that 𝛗nL(Ω)R\|{\bm{\varphi}}_{n}\|_{L^{\infty}(\Omega)}\leq R for all nn\in{\mathbb{N}} and 𝛗n𝛗{\bm{\varphi}}_{n}\to{\bm{\varphi}}_{*} strongly in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}) as nn\to\infty. Let (𝐮¯n,𝐮^n)({\overline{{\bf u}}}_{n},{\widehat{{\bf u}}}_{n}) denote the solutions to (3.9) and (3.10) corresponding to data 𝛗n{\bm{\varphi}}_{n}, 𝐅¯\overline{{\bf F}}, 𝐅^\widehat{{\bf F}}, 𝐠¯\overline{{\bf g}}, 𝐠^\widehat{{\bf g}}, H¯{\overline{H}} and H^{\widehat{H}}. Then, it holds that, as nn\to\infty,

𝐮¯n𝐮¯ strongly in H¯D1(Ω,d),𝐮^n𝐮^ strongly in H^D1(Ω,d),{\overline{{\bf u}}}_{n}\to{\overline{{\bf u}}}_{*}\text{ strongly in }\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\quad{\widehat{{\bf u}}}_{n}\to{\widehat{{\bf u}}}_{*}\text{ strongly in }\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),

where (𝐮¯,𝐮^)H¯D1(Ω,d)×H^D1(Ω,d)({\overline{{\bf u}}}_{*},{\widehat{{\bf u}}}_{*})\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) are the unique solutions to (3.9) and (3.10) corresponding to data 𝛗{\bm{\varphi}}_{*}, 𝐅¯\overline{{\bf F}}, 𝐅^\widehat{{\bf F}}, 𝐠¯\overline{{\bf g}}, 𝐠^\widehat{{\bf g}}, H¯{\overline{H}} and H^{\widehat{H}}.

Proof.

From (3.11) we infer that 𝐮¯n{\overline{{\bf u}}}_{n} and 𝐮^n{\widehat{{\bf u}}}_{n} are bounded in H¯D1(Ω,d)\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), respectively, and thus there exist limit functions (𝐮¯,𝐮^)H¯D1(Ω,d)×H^D1(Ω,d)({\overline{{\bf u}}}_{*},{\widehat{{\bf u}}}_{*})\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) such that, along a non-relabelled subsequence, 𝐮¯n𝐮¯{\overline{{\bf u}}}_{n}\rightharpoonup{\overline{{\bf u}}}_{*} in H¯D1(Ω,d)\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝐮^n𝐮^{\widehat{{\bf u}}}_{n}\rightharpoonup{\widehat{{\bf u}}}_{*} in H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). To obtain strong convergence, in (3.13) and (3.14) we substitute 𝝋1=𝝋n{\bm{\varphi}}_{1}={\bm{\varphi}}_{n}, 𝝋2=𝝋{\bm{\varphi}}_{2}={\bm{\varphi}}_{*}, 𝐮¯1=𝐮¯n{\overline{{\bf u}}}_{1}={\overline{{\bf u}}}_{n}, 𝐮¯2=𝐮¯{\overline{{\bf u}}}_{2}={\overline{{\bf u}}}_{*}, 𝐮¯=𝐮¯n𝐮¯{\overline{{\bf u}}}={\overline{{\bf u}}}_{n}-{\overline{{\bf u}}}_{*} and 𝐮^=𝐮^n𝐮^{\widehat{{\bf u}}}={\widehat{{\bf u}}}_{n}-{\widehat{{\bf u}}}_{*}. Then, by virtue of the dominated convergence theorem, we infer the strong convergence, as nn\to\infty,

(¯(𝝋n)¯(𝝋))(𝐮¯)𝟎 in L2(Ω,d×d).\displaystyle({\overline{{\mathbb{C}}}}({\bm{\varphi}}_{n})-{\overline{{\mathbb{C}}}}({\bm{\varphi}}_{*}))\mathcal{E}({\overline{{\bf u}}}_{*})\to\bm{0}\quad\text{ in }L^{2}(\Omega,{\mathbb{R}}^{d\times d}).

Hence, in the analogue of the first inequality in (3.15) we see that the integral on the right-hand side converges to zero, which implies via Korn’s inequality that

𝐮¯n𝐮¯H1(Ω)0.\mathopen{\|}{\overline{{\bf u}}}_{n}-{\overline{{\bf u}}}_{*}\mathclose{\|}_{H^{1}(\Omega)}\to 0.

By the generalised dominated convergence theorem, we have the strong convergences

{(^(𝝋n)^(𝝋))χ(𝝋n)(𝐮¯n)𝟎^(𝝋)(χ(𝝋n)χ(𝝋))(𝐮^n)𝟎^(𝝋)χ(𝝋)(𝐮¯n𝐮¯)𝟎 in L2(Ω,d×d),\displaystyle\begin{cases}({\widehat{{\mathbb{C}}}}({\bm{\varphi}}_{n})-{\widehat{{\mathbb{C}}}}({\bm{\varphi}}_{*}))\chi({\bm{\varphi}}_{n})\mathcal{E}({\overline{{\bf u}}}_{n})&\to\bm{0}\\ {\widehat{{\mathbb{C}}}}({\bm{\varphi}}_{*})(\chi({\bm{\varphi}}_{n})-\chi({\bm{\varphi}}_{*}))\mathcal{E}({\widehat{{\bf u}}}_{n})&\to\bm{0}\\ \widehat{{\mathbb{C}}}({\bm{\varphi}}_{*})\chi({\bm{\varphi}}_{*})\mathcal{E}({\overline{{\bf u}}}_{n}-{\overline{{\bf u}}}_{*})&\to\bm{0}\end{cases}\text{ in }L^{2}(\Omega,{\mathbb{R}}^{d\times d}),

and thus, in the analogue of the first inequality in (3.17) we see that the integral on the right-hand side converges to zero, leading to the assertion

𝐮^n𝐮^H1(Ω)0.\mathopen{\|}{\widehat{{\bf u}}}_{n}-{\widehat{{\bf u}}}_{*}\mathclose{\|}_{H^{1}(\Omega)}\to 0.

Thus, by combining the weak convergences with the above norms convergence the claim follows. ∎

The above analysis for systems (3.7) and (3.8) allows us to define some solution operators. Namely, we introduce

𝒮:L(Ω,L)H^D1(Ω,d),𝒮:𝝋𝐮^=𝐮^(𝝋),\displaystyle\mathcal{S}:L^{\infty}(\Omega,{\mathbb{R}}^{L})\to\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\quad\mathcal{S}:{\bm{\varphi}}\mapsto{\widehat{{\bf u}}}={\widehat{{\bf u}}}({\bm{\varphi}}), (3.18)

as well as the intermediate operators:

𝒮1:L(Ω,L)L(Ω,L)×H¯D1(Ω,d),\displaystyle\mathcal{S}_{1}:L^{\infty}(\Omega,{\mathbb{R}}^{L})\to L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\quad 𝒮1:𝝋(𝒮11(𝝋),𝒮12(𝝋))=(𝝋,𝐮¯),\displaystyle\mathcal{S}_{1}:{\bm{\varphi}}\mapsto(\mathcal{S}_{1}^{1}({\bm{\varphi}}),\mathcal{S}_{1}^{2}({\bm{\varphi}}))=({\bm{\varphi}},{{\overline{{\bf u}}}}),
𝒮2:L(Ω,L)×H¯D1(Ω,d)H^D1(Ω,d),\displaystyle\mathcal{S}_{2}:L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\to\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\quad 𝒮2:(𝝋,𝐮¯)𝐮^,\displaystyle\mathcal{S}_{2}:({\bm{\varphi}},{\overline{{\bf u}}})\mapsto{\widehat{{\bf u}}},

where 𝐮¯=𝐮¯(𝝋){\overline{{\bf u}}}={\overline{{\bf u}}}({\bm{\varphi}}) and 𝐮^=𝐮^(𝝋,𝐮¯){\widehat{{\bf u}}}={\widehat{{\bf u}}}({\bm{\varphi}},{\overline{{\bf u}}}) are the unique solutions obtained from Theorem 3.1. Then, the overall solution operator 𝒮\mathcal{S} in (3.18) is simply the composition of the intermediate mappings 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2}, i.e., 𝒮=𝒮2𝒮1\mathcal{S}=\mathcal{S}_{2}\circ\mathcal{S}_{1}. In particular, we can define the reduced cost functional

Jred:𝒰ad,Jred:𝝋J(𝝋,𝒮(𝝋)).J_{\rm red}:\mathcal{U}_{\rm ad}\to{\mathbb{R}},\quad J_{\rm red}:{\bm{\varphi}}\mapsto J({\bm{\varphi}},\mathcal{S}({\bm{\varphi}})).

3.3 Existence of optimal designs

Theorem 3.3.

Under (A1)(A6), the optimisation problem (P) admits at least one solution.

Proof.

As the proof is nowadays standard with the direct method of the calculus of variations, let us briefly outline the main points. Consider a minimising sequence {𝝋n}n𝒰ad\{{\bm{\varphi}}_{n}\}_{n\in{\mathbb{N}}}\subset\mathcal{U}_{\rm ad} for the reduced cost functional JredJ_{\rm red}, which satisfies

limnJred(𝝋n)=inf𝝋𝒰adJred(𝝋)0.\displaystyle\lim_{n\to\infty}J_{\rm red}({\bm{\varphi}}_{n})=\inf_{{\bm{\varphi}}\in\mathcal{U}_{\rm ad}}J_{\rm red}({\bm{\varphi}})\geq 0.

This yields that {𝝋n}n\{{\bm{\varphi}}_{n}\}_{n\in{\mathbb{N}}} is bounded in H1(Ω,L)L(Ω,L)H^{1}(\Omega,{\mathbb{R}}^{L})\cap L^{\infty}(\Omega,{\mathbb{R}}^{L}). By standard compactness arguments, since 𝒰ad\mathcal{U}_{\rm ad} is closed and convex, we obtain a limit function 𝝋𝒰ad{\bm{\varphi}}^{*}\in\mathcal{U}_{\rm ad} such that 𝝋n𝝋{\bm{\varphi}}_{n}\to{\bm{\varphi}}^{*} weakly* in H1(Ω,L)L(Ω,L)H^{1}(\Omega,{\mathbb{R}}^{L})\cap L^{\infty}(\Omega,{\mathbb{R}}^{L}) along a non-relabelled subsequence. Consequently, by (3.11) the sequence {𝐮^n=𝒮(𝝋n)}n\{{\widehat{{\bf u}}}_{n}=\mathcal{S}({\bm{\varphi}}_{n})\}_{n\in{\mathbb{N}}} is bounded in H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), and on invoking Corollary 3.1 there exists a limit function 𝐮^H^D1(Ω,d){\widehat{{\bf u}}}^{*}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) such that, along a non-relabelled subsequence, 𝐮^n𝐮^{\widehat{{\bf u}}}_{n}\to{\widehat{{\bf u}}}^{*} strongly in H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) as nn\to\infty. Continuity of the boundary trace operator gives 𝐮^n𝐮^{\widehat{{\bf u}}}_{n}\to{\widehat{{\bf u}}}^{*} strongly in L2(Γ^N,d)L^{2}(\widehat{\Gamma}_{N},{\mathbb{R}}^{d}), and thus

ΓtarW(𝐮^n𝒖tar)(𝐮^n𝒖tar)dd1ΓtarW(𝐮^𝒖tar)(𝐮^𝒖tar)dd1.\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{n}-\bm{u}^{\rm tar})\cdot({\widehat{{\bf u}}}_{n}-\bm{u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1}\to\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{*}-\bm{u}^{\rm tar})\cdot({\widehat{{\bf u}}}^{*}-\bm{u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1}.

By Fatou’s lemma and the a.e. convergence of 𝝋n{\bm{\varphi}}_{n} to 𝝋{\bm{\varphi}}^{*}, we have

lim infnΨ(𝝋n)L1(Ω)Ψ(𝝋)L1(Ω),\displaystyle\liminf_{n\to\infty}\|\Psi({\bm{\varphi}}_{n})\|_{L^{1}(\Omega)}\geq\|\Psi({\bm{\varphi}}^{*})\|_{L^{1}(\Omega)},

and using also the weak lower semicontinuity of the L2L^{2}-norm, we infer that

Jred(𝝋)limnJred(𝝋n)=inf𝝋𝒰adJred(𝝋).J_{\rm red}({\bm{\varphi}}^{*})\leq\lim_{n\to\infty}J_{\rm red}({\bm{\varphi}}_{n})=\inf_{{\bm{\varphi}}\in\mathcal{U}_{\rm ad}}J_{\rm red}({\bm{\varphi}}).

This shows that 𝝋{\bm{\varphi}}^{*} is a solution to (𝐏)(\bf P). ∎

4 Optimality conditions

To derive the first order necessary optimality conditions for 𝝋{\bm{\varphi}}^{*}, we first study the linearised system for the linearised variables introduced below, and use adjoint variables to provide a simplification of the optimality condition.

4.1 Linearised systems and Fréchet differentiability

Here, we analyse some differentiability properties of the solutions operators 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} introduced above. This will help us to formulate the first order optimality conditions of (P).

Theorem 4.1.

The solution operator 𝒮1\mathcal{S}_{1} is Fréchet differentiable at 𝛗{\bm{\varphi}} as a mapping from L(Ω,L)L^{\infty}(\Omega,{\mathbb{R}}^{L}) into L(Ω,L)×H¯D1(Ω,d)L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Moreover, it holds that

D𝒮1(L(Ω,L),L(Ω,L)×H¯D1(Ω,d)),\displaystyle D\mathcal{S}_{1}\in{\cal L}(L^{\infty}(\Omega,{\mathbb{R}}^{L}),L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})),

and its directional derivative at 𝛗L(Ω,L){\bm{\varphi}}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}) along a direction 𝐡L(Ω,L){\bf h}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}) is given by

D𝒮1(𝝋)(𝐡)=(𝐡,𝐯¯),\displaystyle D\mathcal{S}_{1}({\bm{\varphi}})({\bf h})=({\bf h},{\overline{{\bf v}}}), (4.1)

where 𝐯¯H¯D1(Ω,d){\overline{{\bf v}}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) is the unique weak solution to the following system:

div(¯(𝝋)𝐡(𝐮¯)+¯(𝝋)(𝐯¯))\displaystyle-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})+\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf v}}})\big{)} =𝟎\displaystyle=\bm{0}\qquad in Ω,\displaystyle\text{ in }\Omega, (4.2a)
𝐯¯\displaystyle{\overline{{\bf v}}} =𝟎\displaystyle=\bm{0}\qquad on Γ¯D,\displaystyle\text{ on }\overline{\Gamma}_{D}, (4.2b)
(¯(𝝋)𝐡(𝐮¯)+¯(𝝋)(𝐯¯))𝐧\displaystyle(\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})+\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf v}}})){\bf n} =𝟎\displaystyle=\bm{0}\qquad on Γ¯N,\displaystyle\text{ on }\overline{\Gamma}_{N}, (4.2c)

in the sense

Ω¯(𝝋)(𝐯¯):(𝜻)dx+Ω¯(𝝋)𝐡(𝐮¯):(𝜻)dx=0\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\overline{{\bf v}}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}+\int_{\Omega}\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0 (4.3)

for all 𝛇H¯D1(Ω,d){\bm{\zeta}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), where 𝐮¯{\overline{{\bf u}}} is the unique solution to (3.9) associated to 𝛗{\bm{\varphi}} obtained from Theorem 3.1.

Proof.

Firstly, the unique solvability of the linearised system (4.2) follows directly from the application of Proposition 3.1 upon choosing

𝐮=𝐯¯,=¯(𝝋),𝔽=¯(𝝋)𝐡(𝐮¯),𝐟=𝐠=𝟎,ΓD=Γ¯D,ΓN=Γ¯N.\displaystyle{\bf u}={\overline{{\bf v}}},\quad{\mathbb{C}}=\overline{{\mathbb{C}}}({\bm{\varphi}}),\quad{\mathbb{F}}=\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}}),\quad{\bf f}={\bf g}=\bm{0},\quad\Gamma_{D}=\overline{\Gamma}_{D},\quad\Gamma_{N}=\overline{\Gamma}_{N}.

Next, we take 𝝋𝒰ad{\bm{\varphi}}\in\mathcal{U}_{\rm ad} and 𝐡L(Ω,L){\bf h}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}) such that 𝝋𝐡:=𝝋+𝐡𝒰ad{\bm{\varphi}}^{{\bf h}}:={\bm{\varphi}}+{\bf h}\in\mathcal{U}_{\rm ad}, and set 𝐮¯𝐡=𝒮12(𝝋𝐡){\overline{{\bf u}}}^{\bf h}=\mathcal{S}_{1}^{2}({\bm{\varphi}}^{\bf h}), i.e., (𝝋𝐡,𝐮¯𝐡)=𝒮1(𝝋𝐡)({\bm{\varphi}}^{\bf h},{\overline{{\bf u}}}^{\bf h})=\mathcal{S}_{1}({\bm{\varphi}}^{\bf h}). Since the first component of 𝒮1\mathcal{S}_{1} is just the identity in L(Ω,L)L^{\infty}(\Omega,{\mathbb{R}}^{L}), we only need to investigate the Fréchet differentiability of the second component 𝒮12\mathcal{S}_{1}^{2}. In this direction, we denote

𝐰:=𝐮¯𝐡𝐮¯𝐯¯H¯D1(Ω,d)\displaystyle{\bf w}:={\overline{{\bf u}}}^{\bf h}-{\overline{{\bf u}}}-{\overline{{\bf v}}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})

with 𝐯¯{\overline{{\bf v}}} being the unique solution to the linearised system (4.2) associated to 𝝋{\bm{\varphi}} and 𝐡{\bf h}. Our aim is to show the existence of a positive constant CC such that

𝐰H1(Ω)=𝒮12(𝝋𝐡)𝒮12(𝝋)D𝒮12(𝝋)𝐡H1(Ω)C𝐡L(Ω)2,\displaystyle\|{\bf w}\|_{H^{1}(\Omega)}=\|\mathcal{S}_{1}^{2}({\bm{\varphi}}^{\bf h})-\mathcal{S}_{1}^{2}({\bm{\varphi}})-D\mathcal{S}_{1}^{2}({\bm{\varphi}}){\bf h}\|_{H^{1}(\Omega)}\leq C\|{\bf h}\|_{L^{\infty}(\Omega)}^{2}, (4.4)

which would then imply the Fréchet differentiability of the operator 𝒮12\mathcal{S}_{1}^{2}. To this end, we subtract from (3.9) for 𝝋𝐡{\bm{\varphi}}^{\bf h} the sum of (3.9) for 𝝋{\bm{\varphi}} and (4.3) for 𝐯¯{\overline{{\bf v}}} to obtain

Ω¯(𝝋)(𝐰):(𝜻)dx+Ω(¯(𝝋+𝐡)¯(𝝋))((𝐮¯𝐡)(𝐮¯)):(𝜻)dx\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\bf w}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}+\int_{\Omega}(\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\overline{{\bf u}}}^{\bf h})-\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}
+Ω[¯(𝝋+𝐡)¯(𝝋)¯(𝝋)𝐡](𝐮¯):(𝜻)dx=0𝜻H¯D1(Ω,d).\displaystyle\quad+\int_{\Omega}[\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}})-\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0\quad\forall{\bm{\zeta}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}).

Choosing 𝜻=𝐰{\bm{\zeta}}={\bf w} and using (2.4) we infer that

C0(𝐰)L2(Ω)2\displaystyle C_{0}\mathopen{\|}\mathcal{E}({\bf w})\mathclose{\|}^{2}_{L^{2}(\Omega)} Ω(¯(𝝋+𝐡)¯(𝝋))((𝐮¯𝐡)(𝐮¯)):(𝐰)dx\displaystyle\leq-\int_{\Omega}(\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\overline{{\bf u}}}^{\bf h})-\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bf w})\,\mathrm{dx}
Ω[¯(𝝋+𝐡)¯(𝝋)¯(𝝋)𝐡](𝐮¯):(𝐰)dx.\displaystyle\quad-\int_{\Omega}[\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}})-\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\bf w})\,\mathrm{dx}.

Lipschitz continuity of ¯\overline{{\mathbb{C}}} yields a positive constant C¯C_{\overline{{\mathbb{C}}}^{\prime}} such that

|¯(𝝋+𝐡)¯(𝝋)¯(𝝋)𝐡||𝐡|01|¯(𝝋+s𝐡)¯(𝝋)|dsC¯|𝐡|2,\displaystyle|\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}})-\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}|\leq|{\bf h}|\int_{0}^{1}|\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}+s{\bf h})-\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}})|\,\mathrm{d}s\leq C_{\overline{{\mathbb{C}}}^{\prime}}|{\bf h}|^{2}, (4.5)

and keeping in mind the estimate obtained from Theorem 3.2:

𝐮¯𝐡𝐮¯H1(Ω)C𝐡L(Ω),\displaystyle\mathopen{\|}{\overline{{\bf u}}}^{\bf h}-{\overline{{\bf u}}}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}, (4.6)

we find that

Ω(¯(𝝋+𝐡)¯(𝝋))((𝐮¯𝐡)(𝐮¯)):(𝐰)dx\displaystyle\int_{\Omega}(\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\overline{{\bf u}}}^{\bf h})-\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bf w})\,\mathrm{dx}
C𝐡L(Ω)(𝐮¯𝐡)(𝐮¯)L2(Ω)(𝐰)L2(Ω)C04(𝐰)L2(Ω)2+C𝐡L(Ω)4,\displaystyle\quad\leq C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}\mathopen{\|}\mathcal{E}({\overline{{\bf u}}}^{\bf h})-\mathcal{E}({\overline{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}\mathopen{\|}\mathcal{E}({\bf w})\mathclose{\|}_{L^{2}(\Omega)}\leq\frac{C_{0}}{4}\mathopen{\|}\mathcal{E}({\bf w})\mathclose{\|}_{L^{2}(\Omega)}^{2}+C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}^{4},
Ω[¯(𝝋+𝐡)¯(𝝋)¯(𝝋)𝐡](𝐮¯):(𝐰)dx\displaystyle\int_{\Omega}[\overline{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\overline{{\mathbb{C}}}({\bm{\varphi}})-\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\bf w})\,\mathrm{dx}
C𝐡L(Ω)2(𝐮¯)L2(Ω)(𝐰)L2(Ω)C04(𝐰)L2(Ω)2+C𝐡L(Ω)4.\displaystyle\quad\leq C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}^{2}\mathopen{\|}\mathcal{E}({\overline{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}\mathopen{\|}\mathcal{E}({\bf w})\mathclose{\|}_{L^{2}(\Omega)}\leq\frac{C_{0}}{4}\mathopen{\|}\mathcal{E}({\bf w})\mathclose{\|}_{L^{2}(\Omega)}^{2}+C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}^{4}.

Then, by Korn’s inequality we infer (4.4), and whence the claimed Frechét differentiability of 𝒮1\mathcal{S}_{1}. ∎

Before presenting the Fréchet differentiability of 𝒮2\mathcal{S}_{2}, let us provide a formal discussion. Recall that 𝐮^=𝒮2(𝝋,𝐮¯){\widehat{{\bf u}}}=\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}}) and thus the directional derivative D𝒮2(𝝋,𝐮¯)(𝐡,𝐤)D\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h},{\bf k}) of 𝒮2\mathcal{S}_{2} at (𝝋,𝐮¯)({\bm{\varphi}},{\overline{{\bf u}}}) along a direction (𝐡,𝐤)({\bf h},{\bf k}) will be given by

D𝒮2(𝝋,𝐮¯)(𝐡,𝐤)=D𝝋𝒮2(𝝋,𝐮¯)(𝐡)+D𝐮¯𝒮2(𝝋,𝐮¯)(𝐤)D\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h},{\bf k})=D_{{\bm{\varphi}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h})+D_{{\overline{{\bf u}}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf k})

where D𝝋D_{{\bm{\varphi}}} and D𝐮¯D_{{\overline{{\bf u}}}} represent the partial derivatives with respect to 𝝋{\bm{\varphi}} and 𝐮¯{\overline{{\bf u}}}, respectively. Hence, we expect ϑ^:=D𝒮2(𝝋,𝐮¯)(𝐡,𝐤)\widehat{\bm{\vartheta}}:=D\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h},{\bf k}) to be a sum of two functions 𝐯^:=D𝝋𝒮2(𝝋,𝐮¯)(𝐡){\widehat{{\bf v}}}:=D_{{\bm{\varphi}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h}) and 𝐰^:=D𝐮¯𝒮2(𝝋,𝐮¯)(𝐤){\widehat{{\bf w}}}:=D_{{\overline{{\bf u}}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf k}), and the Fréchet differentiability of 𝒮2\mathcal{S}_{2} can be established by demonstrating

𝒮2(𝝋+𝐡,𝐮¯+𝐤)𝒮2(𝝋,𝐮¯)ϑ^H1(Ω)C(𝐡,𝐤)L(Ω)×H1(Ω)2.\displaystyle\|\mathcal{S}_{2}({\bm{\varphi}}+{\bf h},{\overline{{\bf u}}}+{\bf k})-\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})-\widehat{\bm{\vartheta}}\|_{H^{1}(\Omega)}\leq C\|({\bf h},{\bf k})\|_{L^{\infty}(\Omega)\times H^{1}(\Omega)}^{2}. (4.7)

The result is formulated as follows.

Theorem 4.2.

The solution operator 𝒮2\mathcal{S}_{2} is Fréchet differentiable at (𝛗,𝐮¯)({\bm{\varphi}},{\overline{{\bf u}}}) as a mapping from L(Ω,L)×H¯D1(Ω,d)L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) into H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Furthermore,

D𝒮2(L(Ω,L)×H¯D1(Ω,d),H^D1(Ω,d)),\displaystyle D\mathcal{S}_{2}\in{\cal L}(L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})),

and its directional derivative at (𝛗,𝐮¯)L(Ω,L)×H¯D1(Ω,d)({\bm{\varphi}},{\overline{{\bf u}}})\in L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) along a direction (𝐡,𝐤)L(Ω,L)×H¯D1(Ω,d)({\bf h},{\bf k})\in L^{\infty}(\Omega,{\mathbb{R}}^{L})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) is given by

ϑ^:=D𝒮2(𝝋,𝐮¯)(𝐡,𝐤)=D𝝋𝒮2(𝝋,𝐮¯)(𝐡)+D𝐮¯𝒮2(𝝋,𝐮¯)(𝐤)=:𝐯^+𝐰^,\displaystyle\widehat{\bm{\vartheta}}:=D\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h},{\bf k})=D_{\bm{\varphi}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h})+D_{\overline{{\bf u}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf k})=:{\widehat{{\bf v}}}+{\widehat{{\bf w}}},

where ϑ^H^D1(Ω,d)\widehat{\bm{\vartheta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) is the unique weak solution to the following system:

div(^(𝝋)𝐡((𝐮^)χ(𝝋)(𝐮¯)))\displaystyle-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}))\big{)}
div(^(𝝋)((ϑ^)χ(𝝋)𝐡(𝐮¯)χ(𝝋)(𝐤)))\displaystyle\quad-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}(\widehat{\bm{\vartheta}})-\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\bf k}))\big{)} =𝟎\displaystyle=\bm{0}\qquad in Ω,\displaystyle\text{ in }\Omega, (4.8a)
ϑ^\displaystyle\widehat{\bm{\vartheta}} =𝟎\displaystyle=\bm{0}\qquad on Γ^D,\displaystyle\text{ on }\widehat{\Gamma}_{D}, (4.8b)
(^(𝝋)𝐡((𝐮^)χ(𝝋)(𝐮¯)))𝐧\displaystyle\big{(}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}))\big{)}{\bf n}
+(^(𝝋)((ϑ^)χ(𝝋)𝐡(𝐮¯)χ(𝝋)(𝐤)))𝐧\displaystyle\quad+\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}(\widehat{\bm{\vartheta}})-\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\bf k}))\big{)}{\bf n} =𝟎\displaystyle=\bm{0}\qquad on Γ^N,\displaystyle\text{ on }\widehat{\Gamma}_{N}, (4.8c)

in the sense that

Ω^(𝝋)((ϑ^)χ(𝝋)𝐡(𝐮¯)χ(𝝋)(𝐤)):(𝜻)dx\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}(\widehat{\bm{\vartheta}})-\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\bf k})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} (4.9)
+Ω^(𝝋)𝐡((𝐮^)χ(𝝋)(𝐮¯)):(𝜻)dx=0\displaystyle\quad+\int_{\Omega}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0

for all 𝛇H^D1(Ω,d){\bm{\zeta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), where 𝐮¯{\overline{{\bf u}}} is the unique solution to (3.9) associated to 𝛗{\bm{\varphi}} and 𝐮^{\widehat{{\bf u}}} is the unique solution to (3.10) associated to (𝛗,𝐮¯)({\bm{\varphi}},{\overline{{\bf u}}}). Moreover, 𝐯^H^D1(Ω,d){\widehat{{\bf v}}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝐰^H^D1(Ω,d){\widehat{{\bf w}}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) are the unique solutions to the following equations

Ω^(𝝋)((𝐯^)χ(𝝋)𝐡(𝐮¯)):(𝜻)+^(𝝋)𝐡((𝐮^)χ(𝝋)(𝐮¯)):(𝜻)dx=0,\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf v}}})-\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})+\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0, (4.10)
Ω^(𝝋)((𝐰^)χ(𝝋)(𝐤)):(𝜻)dx=0,\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf w}}})-\chi({\bm{\varphi}})\mathcal{E}({\bf k})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0, (4.11)

for all 𝛇H^D1(Ω,d){\bm{\zeta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}).

Proof.

Unique solvability of (4.9), (4.10) and (4.11) are obtained by application of Theorem 3.1, and thus we focus on demonstrating the crucial estimate (4.7). Let 𝝋𝒰ad{\bm{\varphi}}\in\mathcal{U}_{\rm ad}, 𝐡L(Ω,L){\bf h}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}) such that 𝝋𝐡=𝝋+𝐡𝒰ad{\bm{\varphi}}^{\bf h}={\bm{\varphi}}+{\bf h}\in\mathcal{U}_{\rm ad} and 𝐤H¯D1(Ω,d){\bf k}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Setting

𝐮^=𝒮2(𝝋,𝐮¯),𝐮^𝐤:=𝒮2(𝝋,𝐮¯+𝐤),𝐮^𝐡,𝐤:=𝒮2(𝝋+𝐡,𝐮¯+𝐤),{\widehat{{\bf u}}}=\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}}),\quad{\widehat{{\bf u}}}^{{\bf k}}:=\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}}+{\bf k}),\quad{\widehat{{\bf u}}}^{{\bf h},{\bf k}}:=\mathcal{S}_{2}({\bm{\varphi}}+{\bf h},{\overline{{\bf u}}}+{\bf k}),

then, setting 𝝃:=𝐮^𝐡,𝐤𝐮^ϑ^\bm{\xi}:={\widehat{{\bf u}}}^{{\bf h},{\bf k}}-{\widehat{{\bf u}}}-\widehat{\bm{\vartheta}}, (4.7) is equivalent to

𝝃H1(Ω)C(𝐡,𝐤)L(Ω)×H1(Ω)2.\|\bm{\xi}\|_{H^{1}(\Omega)}\leq C\|({\bf h},{\bf k})\|_{L^{\infty}(\Omega)\times H^{1}(\Omega)}^{2}.

We observe that by subtracting from (3.10) for (𝝋+𝐡,𝐮¯+𝐤,𝐮^𝐡,𝐤)({\bm{\varphi}}+{\bf h},{\overline{{\bf u}}}+{\bf k},{\widehat{{\bf u}}}^{{\bf h},{\bf k}}) the sum of (3.10) for (𝝋,𝐮¯,𝐮^)({\bm{\varphi}},{\overline{{\bf u}}},{\widehat{{\bf u}}}) and (4.9) for ϑ^\widehat{\bm{\vartheta}}, we obtain

Ω(^(𝝋+𝐡)((𝐮^𝐡,𝐤)χ(𝝋+𝐡)(𝐮¯+𝐤)):(𝜻)dx\displaystyle\int_{\Omega}\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})(\mathcal{E}({\widehat{{\bf u}}}^{{\bf h},{\bf k}})-\chi({\bm{\varphi}}+{\bf h})\mathcal{E}({\overline{{\bf u}}}+{\bf k})\big{)}:\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} (4.12)
Ω^(𝝋)((𝐮^)χ(𝝋)(𝐮¯)):(𝜻)dx\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}
Ω^(𝝋)𝐡((𝐮^)χ(𝝋)(𝐮¯)):(𝜻)dx\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}(\mathcal{E}({\widehat{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}
Ω^(𝝋)((ϑ^)χ(𝝋)𝐡(𝐮¯)χ(𝝋)(𝐤)):(𝜻)dx=0𝜻H^D1(Ω,d).\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})(\mathcal{E}(\widehat{\bm{\vartheta}})-\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})-\chi({\bm{\varphi}})\mathcal{E}({\bf k})):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=0\quad\forall{\bm{\zeta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}).

Before proceeding with some computations, let us point out the following identities:

^(𝝋+𝐡)(𝐮^𝐡,𝐤)^(𝝋)(𝐮^)^(𝝋)𝐡(𝐮^)^(𝝋)(ϑ^)\displaystyle\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})\mathcal{E}({\widehat{{\bf u}}}^{{\bf h},{\bf k}})-\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}({\widehat{{\bf u}}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\widehat{{\bf u}}})-\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\widehat{\bm{\vartheta}}) (4.13)
=^(𝝋)(𝝃)+(^(𝝋+𝐡)^(𝝋))((𝐮^𝐡,𝐤)(𝐮^𝐡))\displaystyle\quad=\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\bm{\xi})+(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\widehat{{\bf u}}}^{{\bf h},{\bf k}})-\mathcal{E}({\widehat{{\bf u}}}^{\bf h}))
+(^(𝝋+𝐡)^(𝝋))((𝐮^𝐡)(𝐮^))\displaystyle\qquad+(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\widehat{{\bf u}}}^{{\bf h}})-\mathcal{E}({\widehat{{\bf u}}}))
+[^(𝝋+𝐡)^(𝝋)^(𝝋)𝐡](𝐮^)=:^(𝝋)(𝝃)+𝒴1,\displaystyle\qquad+[\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\widehat{{\bf u}}})=:\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\bm{\xi})+{\cal Y}_{1},

and

^(𝝋+𝐡)χ(𝝋+𝐡)(𝐮¯+𝐤)+^(𝝋)χ(𝝋)(𝐮¯)+^(𝝋)𝐡χ(𝝋)(𝐮¯)\displaystyle-\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})\chi({\bm{\varphi}}+{\bf h})\mathcal{E}({\overline{{\bf u}}}+{\bf k})+\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})+\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}}) (4.14)
+^(𝝋)χ(𝝋)𝐡(𝐮¯)+^(𝝋)χ(𝝋)(𝐤)\displaystyle\qquad+\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi^{\prime}({\bm{\varphi}}){\bf h}\mathcal{E}({\overline{{\bf u}}})+\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}({\bf k})
=[^(𝝋+𝐡)^(𝝋)^(𝝋)𝐡]χ(𝝋)(𝐮¯)\displaystyle\quad=-[\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})
^(𝝋)[χ(𝝋+𝐡)χ(𝝋)χ(𝝋)𝐡](𝐮¯)\displaystyle\qquad-\widehat{{\mathbb{C}}}({\bm{\varphi}})[\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}})-\chi^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\overline{{\bf u}}})
(^(𝝋+𝐡)^(𝝋))(χ(𝝋+𝐡)χ(𝝋))(𝐮¯)\displaystyle\qquad-(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}}))\mathcal{E}({\overline{{\bf u}}})
(^(𝝋+𝐡)^(𝝋))(χ(𝝋+𝐡)χ(𝝋))(𝐤)\displaystyle\qquad-(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}}))\mathcal{E}({\bf k})
^(𝝋)(χ(𝝋+𝐡)χ(𝝋))(𝐤)\displaystyle\qquad-\widehat{{\mathbb{C}}}({\bm{\varphi}})(\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}}))\mathcal{E}({\bf k})
(^(𝝋+𝐡)^(𝝋))χ(𝝋)(𝐤)=:𝒴2.\displaystyle\qquad-(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))\chi({\bm{\varphi}})\mathcal{E}({\bf k})=:{\cal Y}_{2}.

Similar to (4.5), we have, for positive constants CχC_{\chi^{\prime}} and C^C_{\widehat{{\mathbb{C}}}^{\prime}}, that

|χ(𝝋+𝐡)χ(𝝋)χ(𝝋)𝐡|\displaystyle|\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}})-\chi^{\prime}({\bm{\varphi}}){\bf h}| |𝐡|01|χ(𝝋+s𝐡)χ(𝝋)|dsCχ|𝐡|2,\displaystyle\leq|{\bf h}|\int_{0}^{1}|\chi^{\prime}({\bm{\varphi}}+s{\bf h})-\chi^{\prime}({\bm{\varphi}})|\,\mathrm{d}s\leq C_{\chi^{\prime}}|{\bf h}|^{2}, (4.15)
|^(𝝋+𝐡)^(𝝋)^(𝝋)𝐡|\displaystyle|\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}| |𝐡|01|^(𝝋+s𝐡)^(𝝋)|dsC^|𝐡|2,\displaystyle\leq|{\bf h}|\int_{0}^{1}|\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}+s{\bf h})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}})|\,\mathrm{d}s\leq C_{\widehat{{\mathbb{C}}}^{\prime}}|{\bf h}|^{2},

and upon choosing 𝜻=𝝃{\bm{\zeta}}=\bm{\xi} in (4.12), we infer that

(𝝃)L2(Ω)C𝒴1L2(Ω)+C𝒴2L2(Ω).\displaystyle\mathopen{\|}\mathcal{E}(\bm{\xi})\mathclose{\|}_{L^{2}(\Omega)}\leq C\|{\cal Y}_{1}\|_{L^{2}(\Omega)}+C\|{\cal Y}_{2}\|_{L^{2}(\Omega)}. (4.16)

Then, employing the Young and Hölder inequalities, (4.6), as well as the stability estimates

𝐮^𝐡,𝐤𝐮^𝐡H1(Ω)C𝐤H1(Ω),𝐮^𝐡𝐮^H1(Ω)C𝐡L(Ω)\mathopen{\|}{\widehat{{\bf u}}}^{{\bf h},{\bf k}}-{\widehat{{\bf u}}}^{{\bf h}}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}{\bf k}\mathclose{\|}_{H^{1}(\Omega)},\quad\mathopen{\|}{\widehat{{\bf u}}}^{{\bf h}}-{\widehat{{\bf u}}}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}

deduced from (3.12), we find that

𝒴1L2(Ω)\displaystyle\|{\cal Y}_{1}\|_{L^{2}(\Omega)} C𝐡L(Ω)(𝐮^𝐡,𝐤𝐮^𝐡H1(Ω)+𝐮^𝐡𝐮^H1(Ω)+𝐡L(Ω)(𝐮^)L2(Ω))\displaystyle\leq C\|{\bf h}\|_{L^{\infty}(\Omega)}\Big{(}\mathopen{\|}{\widehat{{\bf u}}}^{{\bf h},{\bf k}}-{\widehat{{\bf u}}}^{\bf h}\mathclose{\|}_{H^{1}(\Omega)}+\mathopen{\|}{\widehat{{\bf u}}}^{{\bf h}}-{\widehat{{\bf u}}}\mathclose{\|}_{H^{1}(\Omega)}+\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}\mathopen{\|}\mathcal{E}({\widehat{{\bf u}}})\mathclose{\|}_{L^{2}(\Omega)}\Big{)}
C(𝐡,𝐤)L(Ω)×H1(Ω)2,\displaystyle\leq C\mathopen{\|}({\bf h},{\bf k})\mathclose{\|}_{L^{\infty}(\Omega)\times H^{1}(\Omega)}^{2},
𝒴2L2(Ω)\displaystyle\|{\cal Y}_{2}\|_{L^{2}(\Omega)} C𝐡L(Ω)(𝐡L(Ω)(𝐮¯)L2(Ω)+𝐡L(Ω)𝐤H1(Ω)+𝐤H1(Ω))\displaystyle\leq C\|{\bf h}\|_{L^{\infty}(\Omega)}\Big{(}\|{\bf h}\|_{L^{\infty}(\Omega)}\|\mathcal{E}({\overline{{\bf u}}})\|_{L^{2}(\Omega)}+\mathopen{\|}{\bf h}\mathclose{\|}_{L^{\infty}(\Omega)}\|{\bf k}\|_{H^{1}(\Omega)}+\|{\bf k}\|_{H^{1}(\Omega)}\Big{)}
C(𝐡,𝐤)L(Ω)×H1(Ω)2,\displaystyle\leq C\mathopen{\|}({\bf h},{\bf k})\mathclose{\|}_{L^{\infty}(\Omega)\times H^{1}(\Omega)}^{2},

and thus we obtain by Korn’s inequality

𝝃H1(Ω)C(𝐡,𝐤)L(Ω)×H1(Ω)2,\mathopen{\|}\bm{\xi}\mathclose{\|}_{H^{1}(\Omega)}\leq C\mathopen{\|}({\bf h},{\bf k})\mathclose{\|}_{L^{\infty}(\Omega)\times H^{1}(\Omega)}^{2},

which verifies the Fréchet differentiability of 𝒮2\mathcal{S}_{2}. Furthermore, it is clear that by the uniqueness of the solutions, the sum 𝐯^+𝐰^{\widehat{{\bf v}}}+{\widehat{{\bf w}}} is equal to ϑ^\widehat{\bm{\vartheta}}. To establish the identification of the partial derivatives D𝝋𝒮2(𝝋,𝐮¯)(𝐡)=𝐯^D_{{\bm{\varphi}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h})={\widehat{{\bf v}}} and D𝐮¯𝒮2(𝝋,𝐮¯)(𝐤)=𝐰^D_{{\overline{{\bf u}}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf k})={\widehat{{\bf w}}}, we argue as follows: Consider 𝐤=𝟎{\bf k}=\bm{0}, so that from (4.11) we obtain that 𝐰^=𝟎{\widehat{{\bf w}}}=\bm{0} and ϑ^=𝐯^\widehat{\bm{\vartheta}}={\widehat{{\bf v}}}. Then, in (4.12) with 𝐤=𝟎{\bf k}=\bm{0}, we now have for 𝝃=𝐮^𝐡𝐮^𝐯^\bm{\xi}={\widehat{{\bf u}}}^{{\bf h}}-{\widehat{{\bf u}}}-{\widehat{{\bf v}}} the estimate (4.16), where

𝒴1\displaystyle{\cal Y}_{1} =(^(𝝋+𝐡)^(𝝋))((𝐮^𝐡)(𝐮^))+[^(𝝋+𝐡)^(𝝋)^(𝝋)𝐡](𝐮^),\displaystyle=(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\mathcal{E}({\widehat{{\bf u}}}^{{\bf h}})-\mathcal{E}({\widehat{{\bf u}}}))+[\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\widehat{{\bf u}}}),
𝒴2\displaystyle{\cal Y}_{2} =[^(𝝋+𝐡)^(𝝋)^(𝝋)𝐡]χ(𝝋)(𝐮¯)\displaystyle=-[\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}})-\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}){\bf h}]\chi({\bm{\varphi}})\mathcal{E}({\overline{{\bf u}}})
^(𝝋)[χ(𝝋+𝐡)χ(𝝋)χ(𝝋)𝐡](𝐮¯)\displaystyle\quad-\widehat{{\mathbb{C}}}({\bm{\varphi}})[\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}})-\chi^{\prime}({\bm{\varphi}}){\bf h}]\mathcal{E}({\overline{{\bf u}}})
(^(𝝋+𝐡)^(𝝋))(χ(𝝋+𝐡)χ(𝝋))(𝐮¯),\displaystyle\quad-(\widehat{{\mathbb{C}}}({\bm{\varphi}}+{\bf h})-\widehat{{\mathbb{C}}}({\bm{\varphi}}))(\chi({\bm{\varphi}}+{\bf h})-\chi({\bm{\varphi}}))\mathcal{E}({\overline{{\bf u}}}),

where we made use of 𝐮^𝐡,𝟎=𝐮^𝐡{\widehat{{\bf u}}}^{{\bf h},\bm{0}}={\widehat{{\bf u}}}^{{\bf h}}. A short calculation shows that

𝐮^𝐡𝐮^𝐯^H1(Ω)C𝐡L(Ω)2,\|{\widehat{{\bf u}}}^{{\bf h}}-{\widehat{{\bf u}}}-{\widehat{{\bf v}}}\|_{H^{1}(\Omega)}\leq C\|{\bf h}\|_{L^{\infty}(\Omega)}^{2},

which entails that D𝝋𝒮2(𝝋,𝐮¯)(𝐡)=𝐯^D_{{\bm{\varphi}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf h})={\widehat{{\bf v}}}. The other identification D𝐮¯𝒮2(𝝋,𝐮¯)(𝐤)=𝐰^D_{{\overline{{\bf u}}}}\mathcal{S}_{2}({\bm{\varphi}},{\overline{{\bf u}}})({\bf k})={\widehat{{\bf w}}} is in fact more straightforward as 𝒮2(𝝋,)\mathcal{S}_{2}({\bm{\varphi}},\cdot) is a linear operator. This completes the proof. ∎

4.2 Adjoint systems

We now move to the investigation of some adjoint systems which, as typically happens in constrained minimisation problems, will allow us to simplify the formulation of the optimality conditions of (𝐏)(\bf P).

Theorem 4.3.

Under (A1)(A6), for every 𝛗L(Ω,L){\bm{\varphi}}\in L^{\infty}(\Omega,{\mathbb{R}}^{L}), there exists a unique solution 𝐪^H^D1(Ω,d)\widehat{{\bf q}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) to

div(^(𝝋)(𝐪^))\displaystyle-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\widehat{{\bf q}})\big{)} =𝟎\displaystyle=\bm{0}\qquad in Ω,\displaystyle\text{ in }\Omega, (4.17a)
𝐪^\displaystyle\widehat{{\bf q}} =𝟎\displaystyle=\bm{0}\qquad on Γ^D,\displaystyle\text{ on }\widehat{\Gamma}_{D}, (4.17b)
(^(𝝋)(𝐪^))𝐧\displaystyle(\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\widehat{{\bf q}})){\bf n} =W(𝐮^𝐮tar)𝒳Γtar\displaystyle=W({\widehat{{\bf u}}}-{\bf u}^{\rm tar})\mathrm{\mathcal{X}}_{\Gamma^{\rm tar}}\qquad on Γ^N,\displaystyle\text{ on }\widehat{\Gamma}_{N}, (4.17c)

in the sense

Ω^(𝝋)(𝒒^):(𝜻)dx=ΓtarW(𝐮^𝐮tar)𝜻dd1𝜻H^D1(Ω,d).\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\widehat{\bm{q}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}-{\bf u}^{\rm tar})\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}\quad\forall{\bm{\zeta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). (4.18)

Moreover, there exists a unique solution 𝐩¯H¯D1(Ω,d)\overline{{\bf p}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) to

div(¯(𝝋)(𝐩¯)^(𝝋)χ(𝝋)(𝐪^)))\displaystyle-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\overline{{\bf p}})-\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}(\widehat{{\bf q}}))\big{)} =𝟎\displaystyle=\bm{0}\qquad in Ω,\displaystyle\text{ in }\Omega, (4.19a)
𝐩¯\displaystyle\overline{{\bf p}} =𝟎\displaystyle=\bm{0}\qquad on Γ¯D,\displaystyle\text{ on }\overline{\Gamma}_{D}, (4.19b)
(¯(𝝋)(𝐩¯)^(𝝋)χ(𝝋)(𝐪^))𝐧\displaystyle(\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\overline{{\bf p}})-\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}(\widehat{{\bf q}})){\bf n} =𝟎\displaystyle=\bm{0}\qquad on Γ¯N,\displaystyle\text{ on }\overline{\Gamma}_{N}, (4.19c)

in the sense

Ω¯(𝝋)(𝒑¯):(𝜻)dx=Ω^(𝝋)χ(𝝋)(𝒒^):(𝜻)dx𝜻H¯D1(Ω,d),\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}({\bm{\varphi}})\mathcal{E}(\overline{\bm{p}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}})\chi({\bm{\varphi}})\mathcal{E}(\widehat{\bm{q}}):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}\quad\forall{\bm{\zeta}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), (4.20)

where 𝐪^H^D1(Ω,d)\widehat{{\bf q}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) is the unique solution to (4.18).

As the proof of existence and uniqueness is simply an application of Theorem 3.1, we omit the details.

Remark 4.1.

Notice that the adjoint variable 𝐩¯\overline{\bm{p}} to the Stage 1 deformation 𝐮¯{\overline{{\bf u}}} is dependent on the adjoint variable 𝐪^\widehat{\bm{q}} to the Stage 2 deformation 𝐮^{\widehat{{\bf u}}}. This backwards dependence has some parallels with the adjoint systems associated to time-dependent state equations, which have to be solved backwards in time.

4.3 First order optimality conditions

Theorem 4.4.

Under (A1)(A6), let 𝛗𝒰ad{\bm{\varphi}}^{*}\in\mathcal{U}_{\rm ad} be a minimiser to JredJ_{\rm red} with corresponding states 𝐮¯=𝒮12(𝛗){\overline{{\bf u}}}^{*}=\mathcal{S}_{1}^{2}({\bm{\varphi}}^{*}), 𝐮^=𝒮(𝛗){\widehat{{\bf u}}}^{*}=\mathcal{S}({\bm{\varphi}}^{*}), and adjoint variables (𝐩¯,𝐪^)(\overline{\bm{p}},\widehat{\bm{q}}) as unique solutions to (4.18) and (4.20) corresponding to (𝛗,𝐮¯,𝐮^)({\bm{\varphi}}^{*},{\overline{{\bf u}}}^{*},{\widehat{{\bf u}}}^{*}). Then, it necessarily holds that

Ω¯(𝝋)(ϕ𝝋)(𝐮¯):(𝐩¯)dx\displaystyle-\int_{\Omega}\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{*})(\bm{\phi}-{\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*}):\mathcal{E}(\overline{{\bf p}})\,\mathrm{dx}
Ω^(𝝋)(ϕ𝝋)((𝐮^)χ(𝝋)(𝐮¯)):(𝐪^)dx\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{*})(\bm{\phi}-{\bm{\varphi}}^{*})(\mathcal{E}({\widehat{{\bf u}}}^{*})-\chi({\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*})):\mathcal{E}(\widehat{{\bf q}})\,\mathrm{dx}
+Ω^(𝝋)χ(𝝋)(ϕ𝝋)(𝐮¯):(𝐪^)dx\displaystyle\quad+\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}}^{*})\chi^{\prime}({\bm{\varphi}}^{*})(\bm{\phi}-{\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*}):\mathcal{E}(\widehat{{\bf q}})\,\mathrm{dx}
+2γεΩ𝝋(ϕ𝝋)dx+γεΩΨ~,𝝋(𝝋)(ϕ𝝋)dx0ϕ𝒰ad,\displaystyle\quad+2\gamma\varepsilon\int_{\Omega}\nabla{\bm{\varphi}}^{*}\cdot\nabla(\bm{\phi}-{\bm{\varphi}}^{*})\,\mathrm{dx}+\frac{\gamma}{\varepsilon}\int_{\Omega}\widetilde{\Psi}_{,{\bm{\varphi}}}({\bm{\varphi}}^{*})\cdot(\bm{\phi}-{\bm{\varphi}}^{*})\,\mathrm{dx}\geq 0\quad\forall\bm{\phi}\in\mathcal{U}_{\rm ad}, (4.21)

where we set Ψ~,𝛗\widetilde{\Psi}_{,{\bm{\varphi}}} as the vector of partial derivatives of Ψ~\widetilde{\Psi}.

Proof.

As 𝒰ad\mathcal{U}_{\rm ad} is a non-empty, closed and convex set, standard results in optimal control and convex analysis yield that the first order necessary optimality condition for 𝝋{\bm{\varphi}}^{*} is

DJred(𝝋),ϕ𝝋0ϕ𝒰ad,\displaystyle\langle DJ_{\rm red}({\bm{\varphi}}^{*}),\bm{\phi}-{\bm{\varphi}}^{*}\rangle\geq 0\quad\forall\bm{\phi}\in\mathcal{U}_{\rm ad},

which reads, in view of Theorem 4.1 and Theorem 4.2, as

ΓtarW(𝐮^𝐮tar)ϑ^dd1\displaystyle\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{*}-{\bf u}^{\rm tar})\cdot\widehat{\bm{\vartheta}}\,\mathrm{d}\mathcal{H}^{d-1} (4.22)
+2γεΩ𝝋(ϕ𝝋)dx+γεΩΨ~,𝝋(𝝋)(ϕ𝝋)dx0𝝋𝒰ad,\displaystyle\quad+2\gamma\varepsilon\int_{\Omega}\nabla{\bm{\varphi}}^{*}\cdot\nabla(\bm{\phi}-{\bm{\varphi}}^{*})\,\mathrm{dx}+\frac{\gamma}{\varepsilon}\int_{\Omega}\widetilde{\Psi}_{,{\bm{\varphi}}}({\bm{\varphi}}^{*})\cdot(\bm{\phi}-{\bm{\varphi}}^{*})\,\mathrm{dx}\geq 0\quad\forall{\bm{\varphi}}\in\mathcal{U}_{\rm ad},

where ϑ^\widehat{\bm{\vartheta}} is the unique solution to (4.9) with 𝝋=𝝋{\bm{\varphi}}={\bm{\varphi}}^{*}, 𝐮¯=𝐮¯{\overline{{\bf u}}}={\overline{{\bf u}}}^{*}, 𝐮^=𝐮^{\widehat{{\bf u}}}={\widehat{{\bf u}}}^{*}, 𝐡=ϕ𝝋{\bf h}=\bm{\phi}-{\bm{\varphi}}^{*} and 𝐤=𝐯¯{\bf k}={\overline{{\bf v}}} from (4.3) (also with 𝐡=ϕ𝝋{\bf h}=\bm{\phi}-{\bm{\varphi}}^{*}). To simplify the above relation, the procedure is to compare the equalities obtained from (4.3) with 𝜻=𝒑¯{\bm{\zeta}}=\overline{\bm{p}}, (4.9) with 𝐤=𝐯¯{\bf k}={\overline{{\bf v}}} and 𝜻=𝒒^{\bm{\zeta}}=\widehat{\bm{q}}, (4.18) with 𝜻=ϑ^{\bm{\zeta}}=\widehat{\bm{\vartheta}} and (4.20) with 𝜻=𝐯¯{\bm{\zeta}}={\overline{{\bf v}}}. A short calculation then shows

ΓtarW(𝐮^𝐮tar)ϑ^dd1\displaystyle\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{*}-{\bf u}^{\rm tar})\cdot\widehat{\bm{\vartheta}}\,\mathrm{d}\mathcal{H}^{d-1} =Ω¯(𝝋)(𝝋𝝋)(𝐮¯):(𝐩¯)dx\displaystyle=-\int_{\Omega}\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{*})({\bm{\varphi}}-{\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*}):\mathcal{E}(\overline{{\bf p}})\,\mathrm{dx}
Ω^(𝝋)(𝝋𝝋)((𝐮^)χ(𝝋)(𝐮¯)):(𝐪^)dx\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{*})({\bm{\varphi}}-{\bm{\varphi}}^{*})(\mathcal{E}({\widehat{{\bf u}}}^{*})-\chi({\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*})):\mathcal{E}(\widehat{{\bf q}})\,\mathrm{dx}
+Ω^(𝝋)χ(𝝋)(𝝋𝝋)(𝐮¯):(𝐪^)dx,\displaystyle\quad+\int_{\Omega}\widehat{{\mathbb{C}}}({\bm{\varphi}}^{*})\chi^{\prime}({\bm{\varphi}}^{*})({\bm{\varphi}}-{\bm{\varphi}}^{*})\mathcal{E}({\overline{{\bf u}}}^{*}):\mathcal{E}(\widehat{{\bf q}})\,\mathrm{dx},

which allows us to remove the dependence of ϑ^\widehat{\bm{\vartheta}} from (4.22) and leads to (4.21). ∎

5 Sharp interface asymptotics

In this section we study the behaviour of solutions in the sharp interface limit ε0\varepsilon\to 0. For ε>0\varepsilon>0, we denote

Eε(𝝋)\displaystyle E_{\varepsilon}({\bm{\varphi}}) =εΩ|𝝋|2dx+1εΩΨ(𝝋)dx,\displaystyle=\varepsilon\int_{\Omega}|\nabla{\bm{\varphi}}|^{2}\,\mathrm{dx}+\frac{1}{\varepsilon}\int_{\Omega}\Psi({\bm{\varphi}})\,\mathrm{dx},
G(𝝋)\displaystyle G({\bm{\varphi}}) =12ΓtarW(𝒮(𝝋)𝐮tar)(𝒮(𝝋)𝐮tar)dd1,\displaystyle=\frac{1}{2}\int_{\Gamma^{\rm tar}}W(\mathcal{S}({\bm{\varphi}})-{\bf u}^{\rm tar})\cdot(\mathcal{S}({\bm{\varphi}})-{\bf u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1},

where 𝒮\mathcal{S} is the solution operator defined in (3.18), so that the corresponding reduced functional can be expressed as the sum Jredε(𝝋)=G(𝝋)+γEε(𝝋)J_{\rm red}^{\varepsilon}({\bm{\varphi}})=G({\bm{\varphi}})+\gamma E_{\varepsilon}({\bm{\varphi}}). The asymptotic behaviour of solutions can be studied under the framework of Γ\Gamma-convergence. In order to state the result some preparation is needed. A function 𝝋L1(Ω,L){\bm{\varphi}}\in L^{1}(\Omega,{\mathbb{R}}^{L}) is termed a function of bounded variation in Ω\Omega, written as 𝝋BV(Ω,L){\bm{\varphi}}\in\mathrm{BV}(\Omega,{\mathbb{R}}^{L}) if there exists a matrix-valued measure D𝝋D{\bm{\varphi}} of dimension L×dL\times d on Ω\Omega such that

j=1LΩφj(div𝝍)jdx=j=1Li=1dΩψij𝑑Diφj,\sum_{j=1}^{L}\int_{\Omega}\varphi_{j}(\mathop{\rm div}\nolimits{\bm{\psi}})_{j}\,\mathrm{dx}=-\sum_{j=1}^{L}\sum_{i=1}^{d}\int_{\Omega}\psi_{i}^{j}dD_{i}\varphi_{j},

for all 𝝍=(ψij)1id,1jL\bm{\psi}=(\psi_{i}^{j})_{1\leq i\leq d,1\leq j\leq L} where ψijCc1(Ω)\psi_{i}^{j}\in C^{1}_{c}(\Omega). Let T={𝒆1,,𝒆L}T=\{\bm{e}_{1},\dots,\bm{e}_{L}\} where Ψ1(0)=T\Psi^{-1}(0)=T. For 𝝋BV(Ω,T){\bm{\varphi}}\in\mathrm{BV}(\Omega,T) we set, for i{1,,L}i\in\{1,...,L\},

E𝝋i={𝒙Ω:𝝋(𝒙)=𝒆i}E_{\bm{\varphi}}^{i}=\{\bm{x}\in\Omega\,:\,{\bm{\varphi}}(\bm{x})=\bm{e}_{i}\}

and define the essential boundary E𝝋i\partial^{*}E_{{\bm{\varphi}}}^{i} as

E𝝋i={𝒙d:limρ0+|E𝝋iBρ(𝒙)||Bρ(𝒙)|{0,1}}\partial^{*}E_{\bm{\varphi}}^{i}=\Big{\{}\bm{x}\in{\mathbb{R}}^{d}\,:\,\lim_{\rho\to 0^{+}}\frac{|E_{\bm{\varphi}}^{i}\cap B_{\rho}(\bm{x})|}{|B_{\rho}(\bm{x})|}\notin\{0,1\}\Big{\}}

where, for any ρ>0\rho>0, Bρ(𝒙)B_{\rho}(\bm{x}) is the ρ\rho-ball in d{\mathbb{R}}^{d} centered in 𝒙\bm{x}, i.e., Bρ(𝒙)={𝒚d:|𝒚𝒙|<ρ}B_{\rho}(\bm{x})=\{\bm{y}\in{\mathbb{R}}^{d}\,:\,|\bm{y}-\bm{x}|<\rho\}. Consider the extended functionals

𝔼ε(𝝋)\displaystyle{\mathbb{E}}_{\varepsilon}({\bm{\varphi}}) :={Eε(𝝋) if 𝝋H1(Ω,L),+ elsewhere in L1(Ω,L),\displaystyle:=\begin{cases}E_{\varepsilon}({\bm{\varphi}})&\text{ if }{\bm{\varphi}}\in H^{1}(\Omega,{\mathbb{R}}^{L}),\\ +\infty&\text{ elsewhere in }L^{1}(\Omega,{\mathbb{R}}^{L}),\end{cases}
𝔼0(𝝋)\displaystyle{\mathbb{E}}_{0}({\bm{\varphi}}) :={i,j=1,i<jLbijd1(ΩE𝝋iE𝝋j) if 𝝋BV(Ω,T),+ elsewhere in L1(Ω,L),\displaystyle:=\begin{cases}\sum_{i,j=1,\,i<j}^{L}b_{ij}\mathcal{H}^{d-1}(\Omega\cap\partial^{*}E_{{\bm{\varphi}}}^{i}\cap\partial^{*}E_{{\bm{\varphi}}}^{j})&\text{ if }{\bm{\varphi}}\in\mathrm{BV}(\Omega,T),\\ +\infty&\text{ elsewhere in }L^{1}(\Omega,{\mathbb{R}}^{L}),\end{cases}

with constants bijb_{ij} defined as

bij=inf{01Ψ1/2(𝜸(t))|𝜸(t)|dt:𝜸C1([0,1];ΔL),𝜸(0)=𝒆i,𝜸(1)=𝒆j}.b_{ij}=\inf\Big{\{}\int_{0}^{1}\Psi^{1/2}(\bm{\gamma}(t))|\bm{\gamma}^{\prime}(t)|\,\mathrm{d}t\,:\,\bm{\gamma}\in C^{1}([0,1];\Delta^{L}),\,\bm{\gamma}(0)=\bm{e}_{i},\,\bm{\gamma}(1)=\bm{e}_{j}\Big{\}}.

Then, the Γ\Gamma-convergence of 𝔼ε{\mathbb{E}}_{\varepsilon} to 𝔼0{\mathbb{E}}_{0} as ε0\varepsilon\to 0 is expressed as the following assertions:

  • Liminf property. If {𝝋ε}ε>0\{{\bm{\varphi}}^{\varepsilon}\}_{\varepsilon>0} is a sequence such that lim infε0𝔼ε(𝝋ε)<\liminf_{\varepsilon\to 0}\mathbb{E}_{\varepsilon}({\bm{\varphi}}^{\varepsilon})<\infty and 𝝋ε𝝋0{\bm{\varphi}}^{\varepsilon}\to{\bm{\varphi}}^{0} in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}), then 𝝋0BV(Ω,T){\bm{\varphi}}^{0}\in\mathrm{BV}(\Omega,T) with 𝔼0(𝝋0)lim infε0𝔼ε(𝝋ε)\mathbb{E}_{0}({\bm{\varphi}}^{0})\leq\liminf_{\varepsilon\to 0}\mathbb{E}_{\varepsilon}({\bm{\varphi}}^{\varepsilon}).

  • Limsup property. For any 𝝋0L1(Ω,T){\bm{\varphi}}^{0}\in L^{1}(\Omega,T), there exists a sequence {𝝋ε}ε>0H1(Ω,L)\{{\bm{\varphi}}^{\varepsilon}\}_{\varepsilon>0}\subset H^{1}(\Omega,{\mathbb{R}}^{L}), such that 𝝋ε𝝋0{\bm{\varphi}}^{\varepsilon}\to{\bm{\varphi}}^{0} in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}) and lim supε0𝔼ε(𝝋ε)𝔼0(𝝋0)\limsup_{\varepsilon\to 0}{\mathbb{E}}_{\varepsilon}({\bm{\varphi}}^{\varepsilon})\leq\mathbb{E}_{0}({\bm{\varphi}}^{0}).

  • Compactness property. Let {𝝋ε}ε>0\{{\bm{\varphi}}^{\varepsilon}\}_{\varepsilon>0} be a sequence such that supε𝔼ε(𝝋ε)<\sup_{\varepsilon}\mathbb{E}_{\varepsilon}({\bm{\varphi}}^{\varepsilon})<\infty. Then, there exists a non-relabelled subsequence and a function 𝝋0BV(Ω,T){\bm{\varphi}}^{0}\in\mathrm{BV}(\Omega,T) such that 𝝋ε𝝋0{\bm{\varphi}}^{\varepsilon}\to{\bm{\varphi}}^{0} in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}).

For a proof we refer to [8, Thm. 2.5 and Prop. 4.1], see also [11, Thm. 3.1 and Rmk. 3.3] with the choice f(z,X)=|X|2f(z,X)=|X|^{2}.

5.1 Convergences of minimisers

Lemma 5.1.

For each ε(0,1]\varepsilon\in(0,1], let 𝛗ε𝒰ad{\bm{\varphi}}^{\varepsilon}\in\mathcal{U}_{\rm ad} denote a minimiser to the extended reduced cost functional Jredε(𝛗)=G(𝛗)+γ𝔼ε(𝛗)J_{\rm red}^{\varepsilon}({\bm{\varphi}})=G({\bm{\varphi}})+\gamma\mathbb{E}_{\varepsilon}({\bm{\varphi}}). Then, there exists a non-relabelled subsequence ε0\varepsilon\to 0 and a limit function 𝛗0BV(Ω,T){\bm{\varphi}}_{0}\in\mathrm{BV}(\Omega,T) such that 𝛗ε𝛗0{\bm{\varphi}}^{\varepsilon}\to{\bm{\varphi}}_{0} strongly in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}), limε0Jredε(𝛗ε)=Jred0(𝛗0)\lim_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})=J_{\rm red}^{0}({\bm{\varphi}}_{0}), where

Jred0(𝝋)=G(𝝋)+γ𝔼0(𝝋) for 𝝋BV(Ω,T),\displaystyle J_{\rm red}^{0}({\bm{\varphi}})=G({\bm{\varphi}})+\gamma\mathbb{E}_{0}({\bm{\varphi}})\text{ for }{\bm{\varphi}}\in\mathrm{BV}(\Omega,T),

and 𝛗0{\bm{\varphi}}_{0} is a minimiser to Jred0J_{\rm red}^{0}.

Proof.

The proof relies on the Γ\Gamma-convergence of the Ginzburg–Landau functional and the stability of Γ\Gamma-convergence under continuous perturbations. By Corollary 3.1, and the continuity of the trace operator, we see that GG is continuous. For arbitrary 𝝍BV(Ω,T)\bm{\psi}\in\mathrm{BV}(\Omega,T), we invoke the limsup property to find a sequence {𝝍ε}ε>0\{\bm{\psi}^{\varepsilon}\}_{\varepsilon>0} such that 𝝍ε𝝍\bm{\psi}^{\varepsilon}\to\bm{\psi} strongly in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}) and lim supε0𝔼ε(𝝍ε)𝔼0(𝝍)<\limsup_{\varepsilon\to 0}{\mathbb{E}}_{\varepsilon}(\bm{\psi}^{\varepsilon})\leq{\mathbb{E}}_{0}(\bm{\psi})<\infty. Continuity of GG implies G(𝝍ε)G(𝝍)G(\bm{\psi}^{\varepsilon})\to G(\bm{\psi}) as ε0\varepsilon\to 0, leading to

lim supε0Jredε(𝝍ε)Jred0(𝝍)<.\limsup_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}(\bm{\psi}^{\varepsilon})\leq J_{\rm red}^{0}(\bm{\psi})<\infty.

As 𝝋ε{\bm{\varphi}}^{\varepsilon} minimises JredεJ_{\rm red}^{\varepsilon}, we see that

lim supε0Jredε(𝝋ε)lim supε0Jredε(𝝍ε)Jred0(𝝍)<.\limsup_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})\leq\limsup_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}(\bm{\psi}^{\varepsilon})\leq J_{\rm red}^{0}(\bm{\psi})<\infty.

By the non-negativity of GG, the above estimate implies supε(0,1]𝔼ε(𝝋ε)<\sup_{\varepsilon\in(0,1]}{\mathbb{E}}_{\varepsilon}({\bm{\varphi}}^{\varepsilon})<\infty, and by the compactness property we deduce that there exists a limit function 𝝋0BV(Ω,T){\bm{\varphi}}_{0}\in\mathrm{BV}(\Omega,T) such that 𝝋ε𝝋0{\bm{\varphi}}^{\varepsilon}\to{\bm{\varphi}}_{0} strongly in L1(Ω,L)L^{1}(\Omega,{\mathbb{R}}^{L}) along a non-relabelled subsequence. Continuity of GG then gives G(𝝋ε)G(𝝋0)G({\bm{\varphi}}^{\varepsilon})\to G({\bm{\varphi}}_{0}) and invoking the liminf property we subsequently infer that

Jred0(𝝋0)lim infε0Jredε(𝝋ε)Jred0(𝝍).J_{\rm red}^{0}({\bm{\varphi}}_{0})\leq\liminf_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})\leq J_{\rm red}^{0}(\bm{\psi}).

As 𝝍\bm{\psi} is arbitrary, this shows that 𝝋0{\bm{\varphi}}_{0} is a minimiser of Jred0J_{\rm red}^{0}. We now return to the beginning of the proof and consider using the limsup property to construct a sequence {𝝋ε}ε>0\{\bm{\varphi}^{\varepsilon}\}_{\varepsilon>0} that converges strongly to 𝝋0{\bm{\varphi}}_{0} in L1(Ω,d)L^{1}(\Omega,{\mathbb{R}}^{d}). Then, following a similar argument we arrive at

Jred0(𝝋0)lim infε0Jredε(𝝋ε)lim supε0Jredε(𝝋ε)Jred0(𝝋0),J_{\rm red}^{0}({\bm{\varphi}}_{0})\leq\liminf_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})\leq\limsup_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})\leq J_{\rm red}^{0}({\bm{\varphi}}_{0}),

which provides the claimed assertion limε0Jredε(𝝋ε)=Jred0(𝝋0)\lim_{\varepsilon\to 0}J_{\rm red}^{\varepsilon}({\bm{\varphi}}^{\varepsilon})=J_{\rm red}^{0}({\bm{\varphi}}_{0}). ∎

5.2 Formally matched asymptotic expansions

We turn our attention towards the optimality condition (4.21) and study its sharp interface limit ε0\varepsilon\to 0 with the method of formally matched asymptotic expansions, where we assume the functions 𝝋ε{\bm{\varphi}}^{\varepsilon}, 𝐮¯ε{\overline{{\bf u}}}^{\varepsilon}, 𝐮^ε{\widehat{{\bf u}}}^{\varepsilon}, 𝐩¯ε\overline{{\bf p}}^{\varepsilon}, and 𝐪^ε\widehat{{\bf q}}^{\varepsilon} admit asymptotic expansions in powers of ε\varepsilon. From Lemma 5.1 we saw that 𝝋ε{\bm{\varphi}}^{\varepsilon} converges to a function 𝝋0BV(Ω,T){\bm{\varphi}}_{0}\in\mathrm{BV}(\Omega,T) as ε0\varepsilon\to 0, and thus for 0<ε<10<\varepsilon<1, we expect 𝝋ε{\bm{\varphi}}^{\varepsilon} to change its values rapidly on a length scale proportional to ε\varepsilon. This inspires us to consider two asymptotic expansions of (𝝋ε,𝐮¯ε,𝐮^ε,𝐩¯ε,𝐪^ε)({\bm{\varphi}}^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon},\overline{{\bf p}}^{\varepsilon},\widehat{{\bf q}}^{\varepsilon}) in the bulk and interfacial regions (to be defined below), and the procedure is to match these expansions in an intermediate region to deduce the expected equations in the sharp interface limit. We follow the ideas in [13] that treats a similar system of equations, and refer the reader to, e.g., [18, 25, 32, 33] for more details regarding the methodology.

Recalling T=Ψ1(0)={𝒆1,,𝒆L}T=\Psi^{-1}(0)=\{\bm{e}_{1},\dots,\bm{e}_{L}\} as the set of corners of the Gibbs simplex ΔL\Delta^{L}, we partition the domain Ω\Omega into regions Ωi\Omega^{i}, i=1,,Li=1,\dots,L, where Ωi={𝒙Ω:𝝋0(𝒙)=𝒆i}\Omega^{i}=\{\bm{x}\in\Omega\,:\,{\bm{\varphi}}_{0}(\bm{x})=\bm{e}_{i}\}. Then, we assume the functions 𝝋ε{\bm{\varphi}}^{\varepsilon}, 𝐮¯ε{\overline{{\bf u}}}^{\varepsilon}, 𝐮^ε{\widehat{{\bf u}}}^{\varepsilon}, 𝐩¯ε\overline{{\bf p}}^{\varepsilon}, and 𝐪^ε\widehat{{\bf q}}^{\varepsilon} are sufficiently smooth and admit the following asymptotic expansion in ε\varepsilon:

𝝋ε(𝒙)=k=0εk𝝋k(𝒙),𝐮¯ε(𝒙)=k=0εk𝐮¯k(𝒙),𝐮^ε(𝒙)=k=0εk𝐮^k(𝒙),\displaystyle{\bm{\varphi}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}{\bm{\varphi}}_{k}(\bm{x}),\quad{\overline{{\bf u}}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}{\overline{{\bf u}}}_{k}(\bm{x}),\quad{\widehat{{\bf u}}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}{\widehat{{\bf u}}}_{k}(\bm{x}),
𝐩¯ε(𝒙)=k=0εk𝐩¯k(𝒙),𝐪^ε(𝒙)=k=0εk𝐪^k(𝒙),\displaystyle\overline{{\bf p}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\overline{{\bf p}}_{k}(\bm{x}),\quad\widehat{{\bf q}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\widehat{{\bf q}}_{k}(\bm{x}),

for all points 𝒙Ω\bm{x}\in\Omega away from the interfaces Γij=ΩiΩj\Gamma_{ij}=\partial\Omega_{i}\cap\partial\Omega_{j} for i,j{1,,L}i,j\in\{1,\dots,L\}, iji\neq j. This is known as the outer expansion. Furthermore, we assume that

𝝋k(𝒙)TΣL:={𝒗=(v1,,vL)L:i=1Lvi=0},k1,{\bm{\varphi}}_{k}(\bm{x})\in T\Sigma^{L}:=\Big{\{}\bm{v}=(v_{1},\dots,v_{L})\in{\mathbb{R}}^{L}\,:\,\sum_{i=1}^{L}v_{i}=0\Big{\}},\quad k\geq 1,

where TΣLT\Sigma^{L} is the tangent space of the affine hyperplane ΣL={𝒗L:i=1Lvi=1}\Sigma^{L}=\{\bm{v}\in{\mathbb{R}}^{L}\,:\,\sum_{i=1}^{L}v_{i}=1\}, so that by the above construction 𝝋ε(𝒙)ΔL{\bm{\varphi}}^{\varepsilon}(\bm{x})\in\Delta^{L} for ε\varepsilon sufficiently small. We assume that there are constant elasticity tensors ¯i\overline{{\mathbb{C}}}_{i} and ^i\widehat{{\mathbb{C}}}_{i} for i=1,,Li=1,\dots,L, satisfying the standard symmetric conditions and are positive definite. Then, for 𝝋=(φ1,,φL){\bm{\varphi}}=(\varphi_{1},\dots,\varphi_{L}) such that 𝝋(𝒙)ΔL{\bm{\varphi}}(\bm{x})\in\Delta^{L}, we consider

¯(𝝋)=i=1L¯iφi,^(𝝋)=i=1L^iφi.\overline{{\mathbb{C}}}({\bm{\varphi}})=\sum_{i=1}^{L}\overline{{\mathbb{C}}}_{i}\varphi_{i},\quad\widehat{{\mathbb{C}}}({\bm{\varphi}})=\sum_{i=1}^{L}\widehat{{\mathbb{C}}}_{i}\varphi_{i}.

Then, substituting the outer expansions into the state systems (2.1), (2.2) and the adjoint systems (4.17) and (4.19), to leading order we obtain the following equations for i=1,,Li=1,\dots,L:

{div(¯i(𝐮¯0))=𝐅¯ in Ωi,𝐮¯0=𝑼¯ on Γ¯DΩi,(¯i(𝐮¯0))𝐧=𝐠¯ on Γ¯NΩi,\displaystyle\begin{cases}-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}_{i}\mathcal{E}({\overline{{\bf u}}}_{0})\big{)}=\overline{{\bf F}}&\text{ in }\Omega^{i},\\ {\overline{{\bf u}}}_{0}=\overline{\bm{U}}&\text{ on }\overline{\Gamma}_{D}\cap\partial\Omega^{i},\\ (\overline{{\mathbb{C}}}_{i}\mathcal{E}({\overline{{\bf u}}}_{0})){\bf n}=\overline{{\bf g}}&\text{ on }\overline{\Gamma}_{N}\cap\partial\Omega^{i},\end{cases} (5.1)
{div(^i((𝐮^0)χi(𝐮¯0)))=𝐅^ in Ωi,𝐮^0=𝑼^ on Γ^DΩi,(^i((𝐮^0)χi(𝐮¯0)))𝐧=𝐠^ on Γ^NΩi,\displaystyle\begin{cases}-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}_{i}(\mathcal{E}({\widehat{{\bf u}}}_{0})-\chi_{i}\mathcal{E}({\overline{{\bf u}}}_{0}))\big{)}=\widehat{{\bf F}}&\text{ in }\Omega^{i},\\ {\widehat{{\bf u}}}_{0}=\widehat{\bm{U}}&\text{ on }\widehat{\Gamma}_{D}\cap\partial\Omega^{i},\\ (\widehat{{\mathbb{C}}}_{i}(\mathcal{E}({\widehat{{\bf u}}}_{0})-\chi_{i}\mathcal{E}({\overline{{\bf u}}}_{0}))){\bf n}=\widehat{{\bf g}}&\text{ on }\widehat{\Gamma}_{N}\cap\partial\Omega^{i},\end{cases}

where χi:=χ(𝒆i)\chi_{i}:=\chi(\bm{e}_{i}), and

{div(^i(𝐪^0))=𝟎 in Ωi,𝐪^0=𝟎 on Γ^DΩi,(^i(𝐪^0))𝐧=W(𝐮^0𝐮tar)𝒳Γtar on Γ^NΩi,\displaystyle\begin{cases}-\mathop{\rm div}\nolimits\big{(}\widehat{{\mathbb{C}}}_{i}\mathcal{E}(\widehat{{\bf q}}_{0})\big{)}=\bm{0}&\text{ in }\Omega^{i},\\ \widehat{{\bf q}}_{0}=\bm{0}&\text{ on }\widehat{\Gamma}_{D}\cap\partial\Omega^{i},\\ (\widehat{{\mathbb{C}}}_{i}\mathcal{E}(\widehat{{\bf q}}_{0})){\bf n}=W({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\mathrm{\mathcal{X}}_{\Gamma^{\rm tar}}&\text{ on }\widehat{\Gamma}_{N}\cap\partial\Omega^{i},\end{cases} (5.2)
{div(¯i(𝐩¯0)^iχi(𝐪^0)))=𝟎 in Ωi,𝐩¯0=𝟎 on Γ¯DΩi,(¯i(𝐩¯0)^iχi(𝐪^0))𝐧=𝟎 on Γ¯NΩi.\displaystyle\begin{cases}-\mathop{\rm div}\nolimits\big{(}\overline{{\mathbb{C}}}_{i}\mathcal{E}(\overline{{\bf p}}_{0})-\widehat{{\mathbb{C}}}_{i}\chi_{i}\mathcal{E}(\widehat{{\bf q}}_{0}))\big{)}=\bm{0}&\text{ in }\Omega^{i},\\ \overline{{\bf p}}_{0}=\bm{0}&\text{ on }\overline{\Gamma}_{D}\cap\partial\Omega^{i},\\ (\overline{{\mathbb{C}}}_{i}\mathcal{E}(\overline{{\bf p}}_{0})-\widehat{{\mathbb{C}}}_{i}\chi_{i}\mathcal{E}(\widehat{{\bf q}}_{0})){\bf n}=\bm{0}&\text{ on }\overline{\Gamma}_{N}\cap\partial\Omega^{i}.\end{cases}

It then remains to furnish the above with boundary conditions for (𝐮¯0,𝐮^0,𝐩¯0,𝐪^0)({\overline{{\bf u}}}_{0},{\widehat{{\bf u}}}_{0},\overline{{\bf p}}_{0},\widehat{{\bf q}}_{0}) on the interfaces Γij\Gamma_{ij} for i,j{1,,L}i,j\in\{1,\dots,L\}, i<ji<j, which we assume are smooth hypersurfaces that can be obtained in the limit ε0\varepsilon\to 0. These boundary conditions can be inferred with the help of a corresponding inner expansion for (𝝋ε,𝐮¯ε,𝐮^ε,𝐩¯ε,𝐪^ε)({\bm{\varphi}}^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon},\overline{{\bf p}}^{\varepsilon},\widehat{{\bf q}}^{\varepsilon}) in the interfacial regions bordering two bulk regions Ωi\Omega^{i} and Ωj\Omega^{j}. To this end, we focus on one particular interface Γij\Gamma_{ij} and introduce a change of coordinates with the help of a local parameterisation 𝜸:Ud1d\bm{\gamma}:U\subset{\mathbb{R}}^{d-1}\to{\mathbb{R}}^{d} of Γij\Gamma_{ij}, where UU is a spatial parameter domain.

Close to 𝜸(U)\bm{\gamma}(U), consider the signed distance function dd such that d(𝒙)>0d(\bm{x})>0 if 𝒙Ωj\bm{x}\in\Omega^{j} and d(𝒙)<0d(\bm{x})<0 if 𝒙Ωi\bm{x}\in\Omega^{i}, so that the unit normal 𝝂\bm{\nu} to Γij\Gamma_{ij} points from Ωi\Omega^{i} to Ωj\Omega^{j}. Introducing the rescaled signed distance z=dεz=\frac{d}{\varepsilon}, a local parameterization of 𝒙d\bm{x}\in{\mathbb{R}}^{d} close to 𝜸(U){\bm{\gamma}}(U) can be given as

𝒙=𝒢ε(𝒔,z)=𝜸(𝒔)+εz𝝂(𝒔),𝒔Ud1,z.\bm{x}={\cal G}^{\varepsilon}(\bm{s},z)=\bm{\gamma}(\bm{s})+\varepsilon z\bm{\nu}(\bm{s}),\quad\bm{s}\in U\subset{\mathbb{R}}^{d-1},\,z\in{\mathbb{R}}.

This representation allows us to infer the following expansions for gradients, divergences and Laplacians [34]:

xb\displaystyle\nabla_{x}b =1εzb^𝝂+Γijb^+𝒪(ε),divx𝒋=1εz𝒋^𝝂+divΓij𝒋^+𝒪(ε),\displaystyle=\frac{1}{\varepsilon}\partial_{z}\widehat{b}\bm{\nu}+\nabla_{\Gamma_{ij}}\widehat{b}+\mathcal{O}(\varepsilon),\quad\mathop{\rm div}\nolimits_{x}\bm{j}=\frac{1}{\varepsilon}\partial_{z}\widehat{\bm{j}}\cdot\bm{\nu}+\mathop{\rm div}\nolimits_{\Gamma_{ij}}\widehat{\bm{j}}+\mathcal{O}(\varepsilon),
x𝒋\displaystyle\nabla_{x}\bm{j} =1εz𝒋^𝝂+Γij𝒋^+𝒪(ε),Δxb=1ε2zzb^1εκΓijzb^+𝒪(1),\displaystyle=\frac{1}{\varepsilon}\partial_{z}\widehat{\bm{j}}\otimes\bm{\nu}+\nabla_{\Gamma_{ij}}\widehat{\bm{j}}+\mathcal{O}(\varepsilon),\quad\Delta_{x}b=\frac{1}{\varepsilon^{2}}\partial_{zz}\widehat{b}-\frac{1}{\varepsilon}\kappa_{\Gamma_{ij}}\partial_{z}\widehat{b}+\mathcal{O}(1),

for scalar functions b(𝒙)=b^(𝒔(𝒙),z(𝒙))b(\bm{x})=\widehat{b}(\bm{s}(\bm{x}),z(\bm{x})) and vector functions 𝒋(𝒙)=𝒋^(𝒔(𝒙),z(𝒙))\bm{j}(\bm{x})=\widehat{\bm{j}}(\bm{s}(\bm{x}),z(\bm{x})), along with the curvature κΓij\kappa_{\Gamma_{ij}} of Γij\Gamma_{ij}, the surface gradient operator Γij\nabla_{\Gamma_{ij}}, and the surface divergence operator divΓij\mathop{\rm div}\nolimits_{\Gamma_{ij}} on Γij\Gamma_{ij}. Then, for points close by Γij\Gamma_{ij}, we assume an inner expansion of the form

𝝋ε(𝒙)=k=0εk𝚽k(𝒔,z),𝐮¯ε(𝒙)=k=0εk𝑼¯k(𝒔,z),𝐮^ε(𝒙)=k=0εk𝑼^k(𝒔,z),\displaystyle{\bm{\varphi}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\bm{\Phi}_{k}(\bm{s},z),\quad{\overline{{\bf u}}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\overline{\bm{U}}_{k}(\bm{s},z),\quad{\widehat{{\bf u}}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\widehat{\bm{U}}_{k}(\bm{s},z),
𝐩¯ε(𝒙)=k=0εk𝑷¯k(𝒔,z),𝐪^ε(𝒙)=k=0εk𝑸^k(𝒔,z).\displaystyle\overline{{\bf p}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\overline{\bm{P}}_{k}(\bm{s},z),\quad\widehat{{\bf q}}^{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\widehat{\bm{Q}}_{k}(\bm{s},z).

Lastly, we assume in a tubular neighborhood of Γij\Gamma_{ij} the outer expansions and the inner expansions hold simultaneously. Within this intermediate region certain matching conditions relating the outer expansions to the inner expansions must hold. For a scalar function b(𝒙)b(\bm{x}) admitting an outer expansion k=0εkbk(𝒙)\sum_{k=0}^{\infty}\varepsilon^{k}b_{k}(\bm{x}) and an inner expansion k=0εkBk(𝒔,z)\sum_{k=0}^{\infty}\varepsilon^{k}B_{k}(\bm{s},z), it holds that (see [34, Appendix D])

B0(𝒔,z)\displaystyle B_{0}(\bm{s},z) {limδ0b0(𝒙+δ𝝂(𝒙))=:b0+(𝒙) for z+,limδ0b0(𝒙δ𝝂(𝒙))=:b0(𝒙) for z,\displaystyle\to\begin{cases}\lim_{\delta\searrow 0}b_{0}(\bm{x}+\delta\bm{\nu}(\bm{x}))=:b_{0}^{+}(\bm{x})&\text{ for }z\to+\infty,\\ \lim_{\delta\searrow 0}b_{0}(\bm{x}-\delta\bm{\nu}(\bm{x}))=:b_{0}^{-}(\bm{x})&\text{ for }z\to-\infty,\end{cases}
zB0(𝒔,z)\displaystyle\partial_{z}B_{0}(\bm{s},z) 0 as z±,\displaystyle\to 0\text{ as }z\to\pm\infty,
zB1(𝒔,z)\displaystyle\partial_{z}B_{1}(\bm{s},z) {limδ0(b0)(𝒙+δ𝝂(𝒙))𝝂(𝒙)=:b0+𝝂 for z+,limδ0(b0)(𝒙δ𝝂(𝒙))𝝂(𝒙)=:b0𝝂 for z,\displaystyle\to\begin{cases}\lim_{\delta\searrow 0}(\nabla b_{0})(\bm{x}+\delta\bm{\nu}(\bm{x}))\cdot\bm{\nu}(\bm{x})=:\nabla b_{0}^{+}\cdot\bm{\nu}&\text{ for }z\to+\infty,\\ \lim_{\delta\searrow 0}(\nabla b_{0})(\bm{x}-\delta\bm{\nu}(\bm{x}))\cdot\bm{\nu}(\bm{x})=:\nabla b_{0}^{-}\cdot\bm{\nu}&\text{ for }z\to-\infty,\end{cases}

for 𝒙Γij\bm{x}\in\Gamma_{ij}. Consequently, we denote the jump of a quantity bb across Γij\Gamma_{ij} as

[b]+:=limδ0b(𝒙+δ𝝂(𝒙))limδ0b(𝒙δ𝝂(𝒙)) for 𝒙Γij.[b]_{-}^{+}:=\lim_{\delta\searrow 0}b(\bm{x}+\delta\bm{\nu}(\bm{x}))-\lim_{\delta\searrow 0}b(\bm{x}-\delta\bm{\nu}(\bm{x}))\quad\text{ for }\bm{x}\in\Gamma_{ij}.

Note that the above matching conditions also apply to vectorial functions. We introduce the orthogonal projection

𝑷TΣ:LTΣL,𝑷TΣ𝝋=𝝋(1Li=1Lφi)𝟏\bm{P}_{T\Sigma}:{\mathbb{R}}^{L}\to T\Sigma^{L},\quad\bm{P}_{T\Sigma}{\bm{\varphi}}={\bm{\varphi}}-\Big{(}\frac{1}{L}\sum_{i=1}^{L}\varphi_{i}\Big{)}\bm{1}

where 𝟏:=(1,,1)\bm{1}:=(1,\dots,1)^{\top}, so that the optimality condition (4.21) can be expressed in the following strong form

2γεΔ𝝋ε+𝑷TΣ(γεΨ~,𝝋(𝝋ε)+^(𝝋ε)χ(𝝋ε)(𝐮¯ε):(𝐪^ε))\displaystyle-2\gamma\varepsilon\Delta{\bm{\varphi}}^{\varepsilon}+\bm{P}_{T\Sigma}\Big{(}\frac{\gamma}{\varepsilon}\widetilde{\Psi}_{,{\bm{\varphi}}}({\bm{\varphi}}^{\varepsilon})+\widehat{{\mathbb{C}}}({\bm{\varphi}}^{\varepsilon})\chi^{\prime}({\bm{\varphi}}^{\varepsilon})\mathcal{E}({\overline{{\bf u}}}^{\varepsilon}):\mathcal{E}(\widehat{{\bf q}}^{\varepsilon})\Big{)} (5.3)
𝑷TΣ(^(𝝋ε)((𝐮^ε)χ(𝝋ε)(𝐮¯ε)):(𝐪^ε)+¯(𝝋ε)(𝐮¯ε):(𝒑¯ε))=𝟎.\displaystyle\quad-\bm{P}_{T\Sigma}\Big{(}\widehat{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{\varepsilon})(\mathcal{E}({\widehat{{\bf u}}}^{\varepsilon})-\chi({\bm{\varphi}}^{\varepsilon})\mathcal{E}({\overline{{\bf u}}}^{\varepsilon})):\mathcal{E}(\widehat{{\bf q}}^{\varepsilon})+\overline{{\mathbb{C}}}^{\prime}({\bm{\varphi}}^{\varepsilon})\mathcal{E}({\overline{{\bf u}}}^{\varepsilon}):\mathcal{E}(\overline{\bm{p}}^{\varepsilon})\Big{)}=\bm{0}.

We substitute the inner expansions into the equations (2.1a), (2.2a), (4.17a), (4.19a), and (5.3) and collect terms of the same order. Then, we perform some computations to deduce the boundary conditions posed on Γij\Gamma_{ij}. As the subsequent analysis is similar to that performed in [13, 32], we omit most of the straightforward details. In the sequel we use the notation

(𝑩)sym=12(𝑩+𝑩),𝐗:=(z𝐗1𝝂+Γij𝐗0)sym,(\bm{B})^{\mathrm{sym}}=\frac{1}{2}(\bm{B}+\bm{B}^{\top}),\quad\mathcal{E}_{\bf X}:=(\partial_{z}{\bf X}_{1}\otimes\bm{\nu}+\nabla_{\Gamma_{ij}}{\bf X}_{0})^{\mathrm{sym}},

for second order tensors 𝑩\bm{B} and for 𝐗{𝑼¯,𝑼^,𝑷¯,𝑸^}{\bf X}\in\{\overline{\bm{U}},\widehat{\bm{U}},\overline{\bm{P}},\widehat{\bm{Q}}\}.

To leading order 𝒪(1ε2)\mathcal{O}(\frac{1}{\varepsilon^{2}}), equations (2.1a) and (4.17a) yield

z(¯(𝚽0)(z𝑼¯0𝝂)sym𝝂)=𝟎,z(^(𝚽0)(z𝑸^0𝝂)sym𝝂)=𝟎.\partial_{z}\Big{(}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})(\partial_{z}\overline{\bm{U}}_{0}\otimes\bm{\nu})^{\mathrm{sym}}\bm{\nu}\Big{)}=\bm{0},\quad\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})(\partial_{z}\widehat{\bm{Q}}_{0}\otimes\bm{\nu})^{\mathrm{sym}}\bm{\nu}\Big{)}=\bm{0}.

Multiplying by 𝑼¯0\overline{\bm{U}}_{0} and 𝑸^0\widehat{\bm{Q}}_{0}, respectively, integrating over zz\in{\mathbb{R}}, integrating by parts and applying the matching conditions allow us to deduce that z𝑼¯0=z𝑸^0=𝟎\partial_{z}\overline{\bm{U}}_{0}=\partial_{z}\widehat{\bm{Q}}_{0}=\bm{0}, and hence both 𝑼¯0\overline{\bm{U}}_{0} and 𝑸^0\widehat{\bm{Q}}_{0} are constant in zz. Applying matching conditions we infer that

[𝐮¯0]ij=[𝒒^0]ij=𝟎 on Γij.[{\overline{{\bf u}}}_{0}]_{i}^{j}=[\widehat{\bm{q}}_{0}]_{i}^{j}=\bm{0}\quad\text{ on }\Gamma_{ij}.

Then, to leading order 𝒪(1ε2)\mathcal{O}(\frac{1}{\varepsilon^{2}}), equations (2.2a) and (4.19a) yield

z(^(𝚽0)(z𝑼^0𝝂)sym𝝂)=𝟎,z(¯(𝚽0)(z𝑷¯0𝝂)sym𝝂)=𝟎,\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})(\partial_{z}\widehat{\bm{U}}_{0}\otimes\bm{\nu})^{\mathrm{sym}}\bm{\nu}\Big{)}=\bm{0},\quad\partial_{z}\Big{(}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})(\partial_{z}\overline{\bm{P}}_{0}\otimes\bm{\nu})^{\mathrm{sym}}\bm{\nu}\Big{)}=\bm{0},

on account of the fact that z𝑼¯0=z𝑸^0=𝟎\partial_{z}\overline{\bm{U}}_{0}=\partial_{z}\widehat{\bm{Q}}_{0}=\bm{0}. Hence, we also obtain

[𝐮^0]ij=[𝒑¯0]ij=𝟎 on Γij.[{\widehat{{\bf u}}}_{0}]_{i}^{j}=[\overline{\bm{p}}_{0}]_{i}^{j}=\bm{0}\quad\text{ on }\Gamma_{ij}.

To first order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}), we get from (2.1a) and (4.17a) that

z(¯(𝚽0)𝑼¯𝝂)=𝟎,z(^(𝚽0)𝑸^𝝂)=𝟎.\displaystyle\partial_{z}\Big{(}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}\bm{\nu}\Big{)}=\bm{0},\quad\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu}\Big{)}=\bm{0}. (5.4)

Integrating over zz\in{\mathbb{R}} and using the matching conditions yields

[¯(𝐮¯0)𝝂]ij=𝟎,[^(𝒒^0)𝝂]ij=𝟎 on Γij.[\overline{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0})\bm{\nu}]_{i}^{j}=\bm{0},\quad[\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0})\bm{\nu}]_{i}^{j}=\bm{0}\quad\text{ on }\Gamma_{ij}.

Similarly, from equations (2.2a) and (4.19a), we obtain to first order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}) that

z(^(𝚽0)(𝑼^χ(𝚽𝟎)𝑼¯)𝝂)=𝟎,z(¯(𝚽0)𝑷¯𝝂χ(𝚽0)^(𝚽0)𝑸^𝝂)=𝟎.\displaystyle\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})(\mathcal{E}_{\widehat{\bm{U}}}-\chi(\bm{\Phi_{0}})\mathcal{E}_{\overline{\bm{U}}})\bm{\nu}\Big{)}=\bm{0},\quad\partial_{z}\Big{(}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{P}}}\bm{\nu}-\chi(\bm{\Phi}_{0})\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu}\Big{)}=\bm{0}. (5.5)

Integrating over zz\in{\mathbb{R}} and applying the matching conditions, we obtain

[^((𝐮^0)χ(𝝋0)(𝐮¯0))]ij𝝂=𝟎,[¯(𝒑¯0)χ(𝝋0)^(𝒒^0)]ij𝝂=𝟎 on Γij.[\widehat{{\mathbb{C}}}(\mathcal{E}({\widehat{{\bf u}}}_{0})-\chi({\bm{\varphi}}_{0})\mathcal{E}({\overline{{\bf u}}}_{0}))]_{i}^{j}\bm{\nu}=\bm{0},\quad[\overline{{\mathbb{C}}}\mathcal{E}(\overline{\bm{p}}_{0})-\chi({\bm{\varphi}}_{0})\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0})]_{i}^{j}\bm{\nu}=\bm{0}\quad\text{ on }\Gamma_{ij}.

Turning now to the optimality condition (5.3), we use the fact that z𝑼¯0=z𝑼^0=z𝑷¯0=z𝑸^0=𝟎\partial_{z}\overline{\bm{U}}_{0}=\partial_{z}\widehat{\bm{U}}_{0}=\partial_{z}\overline{\bm{P}}_{0}=\partial_{z}\widehat{\bm{Q}}_{0}=\bm{0} to see that the elasticity terms do not contribute to leading order 𝒪(1ε2)\mathcal{O}(\frac{1}{\varepsilon^{2}}) and first order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}). Hence, to first order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}) we obtain from (5.3) that

2zz𝚽0𝑷TΣ(Ψ~,𝝋(𝚽0))=𝟎.\displaystyle 2\partial_{zz}\bm{\Phi}_{0}-\bm{P}_{T\Sigma}\Big{(}\widetilde{\Psi}_{,{\bm{\varphi}}}(\bm{\Phi}_{0})\Big{)}=\bm{0}.

Following [18], 𝚽0\bm{\Phi}_{0} can be chosen independent of ss and as the solution to the above ordinary differential system such that limz𝚽0(z)=𝒆i\lim_{z\to-\infty}\bm{\Phi}_{0}(z)=\bm{e}_{i} and limz+𝚽0(z)=𝒆j\lim_{z\to+\infty}\bm{\Phi}_{0}(z)=\bm{e}_{j}. To the next order 𝒪(1)\mathcal{O}(1), we obtain

2γzz𝚽1+γ𝑷TΣ(Ψ~,𝝋𝝋(𝚽0)𝚽1)+2γκΓijz𝚽0+𝑷TΣ(^(𝚽0)χ(𝚽0)𝑼¯:𝑸^)\displaystyle-2\gamma\partial_{zz}\bm{\Phi}_{1}+\gamma\bm{P}_{T\Sigma}\Big{(}\widetilde{\Psi}_{,{\bm{\varphi}}{\bm{\varphi}}}(\bm{\Phi}_{0})\bm{\Phi}_{1}\Big{)}+2\gamma\kappa_{\Gamma_{ij}}\partial_{z}\bm{\Phi}_{0}+\bm{P}_{T\Sigma}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\chi^{\prime}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}\Big{)} (5.6)
𝑷TΣ(^(𝚽0)(𝑼^χ(𝚽0)𝑼¯):𝑸^+¯(𝚽0)𝑼¯:𝑷¯)=𝟎,\displaystyle\quad-\bm{P}_{T\Sigma}\Big{(}\widehat{{\mathbb{C}}}^{\prime}(\bm{\Phi}_{0})(\mathcal{E}_{\widehat{\bm{U}}}-\chi(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}):\mathcal{E}_{\widehat{\bm{Q}}}+\overline{{\mathbb{C}}}^{\prime}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\overline{\bm{P}}}\Big{)}=\bm{0},

where Ψ~,𝝋𝝋\widetilde{\Psi}_{,{\bm{\varphi}}{\bm{\varphi}}} denotes the Hessian matrix of Ψ~\widetilde{\Psi}, and we have used that z𝚽0TΣL\partial_{z}\bm{\Phi}_{0}\in T\Sigma^{L}. Note that by the fact that z𝐗0=𝟎\partial_{z}{\bf X}_{0}=\bm{0} for 𝐗{𝑼¯,𝑼^,𝑷¯,𝑸^}{\bf X}\in\{\overline{\bm{U}},\widehat{\bm{U}},\overline{\bm{P}},\widehat{\bm{Q}}\}, and by the symmetry of the elasticity tensors ijkl=jikl{\mathbb{C}}_{ijkl}={\mathbb{C}}_{jikl}, we have the relations

z𝐗=(zz𝐗1𝝂)sym,𝐘:z𝐗=((𝐘)𝝂)zz𝐗1,\displaystyle\partial_{z}\mathcal{E}_{\bf X}=(\partial_{zz}{\bf X}_{1}\otimes\bm{\nu})^{\mathrm{sym}},\quad{\mathbb{C}}\mathcal{E}_{\bf Y}:\partial_{z}\mathcal{E}_{\bf X}=(({\mathbb{C}}\mathcal{E}_{\bf Y})\bm{\nu})\cdot\partial_{zz}{\bf X}_{1}, (5.7)

for any 𝐗,𝐘{𝑼¯,𝑼^,𝑷¯,𝑸^}{\bf X},{\bf Y}\in\{\overline{\bm{U}},\widehat{\bm{U}},\overline{\bm{P}},\widehat{\bm{Q}}\}. To obtain a solution 𝚽1\bm{\Phi}_{1}, a solvability condition has to hold, which can be derived by multiplying (5.6) with z𝚽0\partial_{z}\bm{\Phi}_{0} and integrating over zz. Using the relations 𝑷TΣ(z𝚽0)=z𝚽0\bm{P}_{T\Sigma}(\partial_{z}\bm{\Phi}_{0})=\partial_{z}\bm{\Phi}_{0}, 𝑷TΣ(z𝚽1)=z𝚽1\bm{P}_{T\Sigma}(\partial_{z}\bm{\Phi}_{1})=\partial_{z}\bm{\Phi}_{1}, after integrating by parts and applying the matching conditions, we obtain

0\displaystyle 0 =γ(2zz𝚽0+𝑷TΣ(Ψ~,𝝋(𝚽0)))=𝟎z𝚽1dz+2γκΓij|z𝚽0|2𝑑z\displaystyle=\int_{-\infty}^{\infty}\gamma\underbrace{(-2\partial_{zz}\bm{\Phi}_{0}+\bm{P}_{T\Sigma}\Big{(}\widetilde{\Psi}_{,{\bm{\varphi}}}(\bm{\Phi}_{0}))\Big{)}}_{=\bm{0}}\cdot\,\partial_{z}\bm{\Phi}_{1}dz+2\gamma\kappa_{\Gamma_{ij}}\int_{-\infty}^{\infty}|\partial_{z}\bm{\Phi}_{0}|^{2}dz (5.8)
+z[^(𝚽0)χ(𝚽0)]𝑼¯:𝑸^z^(𝚽0)𝑼^:𝑸^z¯(𝚽0)𝑼¯:𝑷¯dz.\displaystyle\quad+\int_{-\infty}^{\infty}\partial_{z}[\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\chi(\bm{\Phi}_{0})]\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}-\partial_{z}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}-\partial_{z}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\overline{\bm{P}}}\;dz.

We employ the identities obtained from (5.4), (5.5) and (5.7) to obtain that

^(𝚽0)𝑼^:z𝑸^+^(𝚽0)𝑼^𝝂zz𝑸^1=𝟎,z(^(𝚽0)𝑸^𝝂)=𝟎,\displaystyle-\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}:\partial_{z}\mathcal{E}_{\widehat{\bm{Q}}}+\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}\bm{\nu}\cdot\partial_{zz}\widehat{\bm{Q}}_{1}=\bm{0},\quad\partial_{z}(\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu})=\bm{0},

as well as z𝝂=𝟎\partial_{z}\bm{\nu}=\bm{0} to see that

z^(𝚽0)𝑼^:𝑸^\displaystyle-\partial_{z}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}
=z^(𝚽0)𝑼^:𝑸^^(𝚽0)𝑼^:z𝑸^^(𝚽0)𝑸^:z𝑼^\displaystyle\quad=-\partial_{z}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}-\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}:\partial_{z}\mathcal{E}_{\widehat{\bm{Q}}}-\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}:\partial_{z}\mathcal{E}_{\widehat{\bm{U}}}
+^(𝚽0)𝑼^𝝂zz𝑸^1+^(𝚽0)𝑸^𝝂zz𝑼^1\displaystyle\qquad+\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{U}}}\bm{\nu}\cdot\partial_{zz}\widehat{\bm{Q}}_{1}+\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu}\cdot\partial_{zz}\widehat{\bm{U}}_{1}
+z(^(𝚽0)𝑸^)𝝂z𝑼^1+z(^(𝚽0)(𝑼^χ(𝚽0)𝑼¯))𝝂z𝑸^1\displaystyle\qquad+\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\Big{)}\bm{\nu}\cdot\partial_{z}\widehat{\bm{U}}_{1}+\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})(\mathcal{E}_{\widehat{\bm{U}}}-\chi(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}})\Big{)}\bm{\nu}\cdot\partial_{z}\widehat{\bm{Q}}_{1}
=z(^(𝚽𝟎)𝑼^:𝑸^^(𝚽𝟎)𝑸^𝝂z𝑼^1^(𝚽𝟎)𝑼^𝝂z𝑸^1))\displaystyle\quad=-\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi_{0}})\mathcal{E}_{\widehat{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}-\widehat{{\mathbb{C}}}(\bm{\Phi_{0}})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu}\cdot\partial_{z}\widehat{\bm{U}}_{1}-\widehat{{\mathbb{C}}}(\bm{\Phi_{0}})\mathcal{E}_{\widehat{\bm{U}}}\bm{\nu}\cdot\partial_{z}\widehat{\bm{Q}}_{1}\big{)}\Big{)}
z(^(𝚽0)χ(𝚽0)𝑼¯)𝝂z𝑸^1.\displaystyle\qquad-\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\chi(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}\Big{)}\bm{\nu}\cdot\partial_{z}\widehat{\bm{Q}}_{1}.

Via a similar calculation we infer

z¯(𝚽0)𝑼¯:𝑷¯\displaystyle-\partial_{z}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\overline{\bm{P}}}
=z(¯(𝚽0)𝑼¯:𝑷¯¯(𝚽0)𝑷¯𝝂z𝑼¯1¯(𝚽0)𝑼¯𝝂z𝑷¯1)\displaystyle\quad=-\partial_{z}\Big{(}\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\overline{\bm{P}}}-\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{P}}}\bm{\nu}\cdot\partial_{z}\overline{\bm{U}}_{1}-\overline{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}\bm{\nu}\cdot\partial_{z}\overline{\bm{P}}_{1}\Big{)}
z(^(𝚽0)χ(𝚽0)𝑸^)𝝂z𝑼¯1,\displaystyle\qquad-\partial_{z}\Big{(}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\chi(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\Big{)}\bm{\nu}\cdot\partial_{z}\overline{\bm{U}}_{1},
z[^(𝚽0)χ(𝚽0)]𝑼¯:𝑸^\displaystyle\partial_{z}\big{[}\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\chi(\bm{\Phi}_{0})\big{]}\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}
=z(χ(𝚽0)^(𝚽0)𝑼¯:𝑸^)\displaystyle\quad=\partial_{z}\Big{(}\chi(\bm{\Phi}_{0})\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}:\mathcal{E}_{\widehat{\bm{Q}}}\Big{)}
χ(𝚽0)^(𝚽0)𝑼¯𝝂zz𝑸^1χ(𝚽0)^(𝚽0)𝑸^𝝂zz𝑼¯1,\displaystyle\qquad-\chi(\bm{\Phi}_{0})\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\overline{\bm{U}}}\bm{\nu}\cdot\partial_{zz}\widehat{\bm{Q}}_{1}-\chi(\bm{\Phi}_{0})\widehat{{\mathbb{C}}}(\bm{\Phi}_{0})\mathcal{E}_{\widehat{\bm{Q}}}\bm{\nu}\cdot\partial_{zz}\overline{\bm{U}}_{1},

and thus, setting bij:=2|z𝚽0|2𝑑zb_{ij}:=\int_{-\infty}^{\infty}2|\partial_{z}\bm{\Phi}_{0}|^{2}dz, and applying matching conditions for 𝐗\mathcal{E}_{\bf X} and z𝐗1\partial_{z}{\bf X}_{1} with 𝐗{𝑼¯,𝑼^,𝑷¯,𝑸^}{\bf X}\in\{\overline{\bm{U}},\widehat{\bm{U}},\overline{\bm{P}},\widehat{\bm{Q}}\}, we obtain from (5.8) the solvability condition

0\displaystyle 0 =γκΓijbij+[χ(𝝋0)^(𝐮¯0):(𝒒^0)]ij[^(𝐮^0):(𝒒^0)]ij[¯(𝐮¯0):(𝒑¯0)]ij\displaystyle=\gamma\kappa_{\Gamma_{ij}}b_{ij}+[\chi({\bm{\varphi}}_{0})\widehat{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0}):\mathcal{E}(\widehat{\bm{q}}_{0})]_{i}^{j}-[\widehat{{\mathbb{C}}}\mathcal{E}({\widehat{{\bf u}}}_{0}):\mathcal{E}(\widehat{\bm{q}}_{0})]_{i}^{j}-[\overline{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0}):\mathcal{E}(\overline{\bm{p}}_{0})]_{i}^{j} (5.9)
+[^(𝒒^0)𝝂(𝐮^0)𝝂+^(𝐮^0)𝝂(𝒒^0)𝝂]ij[χ(𝝋0)^(𝐮¯0)𝝂(𝒒^0)𝝂]ij\displaystyle\quad+[\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0})\bm{\nu}\cdot(\nabla{\widehat{{\bf u}}}_{0})\bm{\nu}+\widehat{{\mathbb{C}}}\mathcal{E}({\widehat{{\bf u}}}_{0})\bm{\nu}\cdot(\nabla\widehat{\bm{q}}_{0})\bm{\nu}]_{i}^{j}-[\chi({\bm{\varphi}}_{0})\widehat{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0})\bm{\nu}\cdot(\nabla\widehat{\bm{q}}_{0})\bm{\nu}]_{i}^{j}
+[¯(𝒑¯0)𝝂(𝐮¯0)𝝂+¯(𝐮¯0)𝝂(𝒑¯0)𝝂]ij[χ(𝝋0)^(𝒒^0)𝝂(𝐮¯0)𝝂]ij\displaystyle\quad+[\overline{{\mathbb{C}}}\mathcal{E}(\overline{\bm{p}}_{0})\bm{\nu}\cdot(\nabla{\overline{{\bf u}}}_{0})\bm{\nu}+\overline{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0})\bm{\nu}\cdot(\nabla\overline{\bm{p}}_{0})\bm{\nu}]_{i}^{j}-[\chi({\bm{\varphi}}_{0})\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0})\bm{\nu}\cdot(\nabla{\overline{{\bf u}}}_{0})\bm{\nu}]_{i}^{j}

that has to hold on Γij\Gamma_{ij}. Thus, the sharp interface limit consists of the equations (5.1) and (5.2) posed in Ωi\Omega^{i}, 1iL1\leq i\leq L, furnished by the boundary conditions (5.9) and

[𝐮^0]ij=𝟎,[𝒑¯0]ij=𝟎,[𝐮¯0]ij=𝟎,[𝒒^0]ij=𝟎,\displaystyle[{\widehat{{\bf u}}}_{0}]_{i}^{j}=\bm{0},\quad[\overline{\bm{p}}_{0}]_{i}^{j}=\bm{0},\quad[{\overline{{\bf u}}}_{0}]_{i}^{j}=\bm{0},\quad[\widehat{\bm{q}}_{0}]_{i}^{j}=\bm{0},
[¯(𝐮¯0)𝝂]ij=𝟎,[^(𝒒^0))𝝂]ij=𝟎,[¯(𝐮¯0)𝝂]ij=𝟎,[^(𝒒^0))𝝂]ij=𝟎\displaystyle[\overline{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0})\bm{\nu}]_{i}^{j}=\bm{0},\quad[\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0}))\bm{\nu}]_{i}^{j}=\bm{0},\quad[\overline{{\mathbb{C}}}\mathcal{E}({\overline{{\bf u}}}_{0})\bm{\nu}]_{i}^{j}=\bm{0},\quad[\widehat{{\mathbb{C}}}\mathcal{E}(\widehat{\bm{q}}_{0}))\bm{\nu}]_{i}^{j}=\bm{0}

on Γij\Gamma_{ij}, 1i<jL1\leq i<j\leq L.

Remark 5.1.

It is possible to consider the sharp interface limit near a triple junction where three regions meet. We refer to [13, 18, 53] for more details regarding the asymptotic analysis around a triple junction.

5.3 Rigorous convergence in the two-phase setting

In the two phase case L=2L=2, since 𝝋=(φ1,φ2)Δ2{\bm{\varphi}}=(\varphi_{1},\varphi_{2})\in\Delta^{2}, we may use the difference φ:=φ2φ1[1,1]\varphi:=\varphi_{2}-\varphi_{1}\in[-1,1] to encode the vector 𝝋=(12(1φ),12(1+φ)){\bm{\varphi}}=(\frac{1}{2}(1-\varphi),\frac{1}{2}(1+\varphi)), so that |𝝋|2=12|φ|2|\nabla{\bm{\varphi}}|^{2}=\frac{1}{2}|\nabla\varphi|^{2}. Hence, the problem (P) can be re-phrased in terms of the scalar function φ\varphi ranging between 1-1 and 11, and it suffices to consider the following

(𝑷ε)minφ𝒰adJredε(φ)\displaystyle(\bm{P}^{\varepsilon})\quad\min_{\varphi\in\mathcal{U}_{\rm ad}}J_{\rm red}^{\varepsilon}(\varphi) =12ΓtarW(𝒮(φ)𝒖tar)(𝒮(φ)𝐮tar)dd1\displaystyle=\frac{1}{2}\int_{\Gamma^{\mathrm{tar}}}W(\mathcal{S}(\varphi)-\bm{u}^{\mathrm{tar}})\cdot(\mathcal{S}(\varphi)-{\bf u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1}
+γΩε2|φ|2+1εΨ(φ)dx,\displaystyle\quad+\gamma\int_{\Omega}\frac{\varepsilon}{2}|\nabla\varphi|^{2}+\frac{1}{\varepsilon}\Psi(\varphi)\,\mathrm{dx},

where 𝒰ad={fH1(Ω):f[1,1] a.e. in Ω}\mathcal{U}_{\rm ad}=\{f\in H^{1}(\Omega)\,:\,f\in[-1,1]\text{ a.e.~{}in }\Omega\}, and, as no confusion can arise, we use the short-hand notations 𝒮(φ)\mathcal{S}(\varphi) and Ψ(φ)\Psi(\varphi) for the functions 𝒮(𝝋)\mathcal{S}({\bm{\varphi}}) and Ψ(𝝋)\Psi({\bm{\varphi}}) evaluated at 𝝋=(12(1φ),12(1+φ)){\bm{\varphi}}=(\frac{1}{2}(1-\varphi),\frac{1}{2}(1+\varphi)). On recalling (A3), we hence assume that

Ψ(s)=Ψ~(s)+I[1,1](s) for s,\Psi(s)=\widetilde{\Psi}(s)+I_{[-1,1]}(s)\text{ for }s\in{\mathbb{R}}, (5.10)

for a Ψ~C1,1()\widetilde{\Psi}\in C^{1,1}({\mathbb{R}}). To simplify the calculations, we consider 𝑼¯=𝑼^=𝟎\overline{\bm{U}}=\widehat{\bm{U}}=\bm{0} (homogeneous Dirichlet data), 𝐅¯,𝐅^H1(Ω,d)\overline{{\bf F}},\widehat{{\bf F}}\in H^{1}(\Omega,{\mathbb{R}}^{d}), 𝐠¯H2(Γ¯N,d)\overline{{\bf g}}\in H^{2}(\overline{\Gamma}_{N},{\mathbb{R}}^{d}) and 𝐠^H2(Γ^N,d)\widehat{{\bf g}}\in H^{2}(\widehat{\Gamma}_{N},{\mathbb{R}}^{d}). We now consider deriving an alternative set of optimal conditions for a minimiser φ𝒰ad\varphi^{*}\in\mathcal{U}_{\rm ad} based on geometric variations. To this end, we consider the following admissible transformations and their corresponding velocity fields.

Definition 5.1.

The space 𝒱ad\mathcal{V}_{\mathrm{ad}} of admissible velocity fields is defined as the set of all 𝐕C0([τ,τ]×Ω¯,d){\bf V}\in C^{0}([-\tau,\tau]\times\overline{\Omega},{\mathbb{R}}^{d}), where τ>0\tau>0 is a fixed small constant, such that it holds:

  • 𝐕(t,)C2(Ω¯,d){\bf V}(t,\cdot)\in C^{2}(\overline{\Omega},{\mathbb{R}}^{d}), and C>0\exists C>0 such that 𝐕(,𝒚)𝐕(,𝒙)C0([τ,τ],d)C𝒙𝒚\|{\bf V}(\cdot,\bm{y})-{\bf V}(\cdot,\bm{x})\|_{C^{0}([-\tau,\tau],{\mathbb{R}}^{d})}\leq C\|\bm{x}-\bm{y}\| for all 𝒙,𝒚Ω¯\bm{x},\bm{y}\in\overline{\Omega};

  • 𝐕(t,𝒙)𝐧(𝒙)=0{\bf V}(t,\bm{x})\cdot{\bf n}(\bm{x})=0 for all 𝒙Ω\bm{x}\in\partial\Omega;

  • 𝐕(t,𝒙)=𝟎{\bf V}(t,\bm{x})=\bm{0} for all 𝒙Γ¯DΓ^D\bm{x}\in\overline{\Gamma}_{D}\cup\widehat{\Gamma}_{D}.

Then, the space 𝒯ad\mathcal{T}_{\mathrm{ad}} of admissible transformations is defined as the set of solutions to the ordinary differential equations

tTt(𝒙)=𝐕(t,Tt(𝒙)),T0(𝒙)=𝒙\partial_{t}T_{t}(\bm{x})={\bf V}(t,T_{t}(\bm{x})),\quad T_{0}(\bm{x})=\bm{x}

with 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}}.

Notice that by the second property it holds Tt(Ω)¯=Ω¯\overline{T_{t}(\Omega)}=\overline{\Omega} for all t[τ,τ]t\in[-\tau,\tau]. Let 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}} be an admissible velocity field with corresponding transformation T𝒯adT\in\mathcal{T}_{\mathrm{ad}}. For φ𝒰ad\varphi\in\mathcal{U}_{\rm ad} we define φt:=φTt1\varphi^{t}:=\varphi\circ T_{t}^{-1}, along with the unique solutions (𝐮¯t,𝐮^t)H¯D1(Ω,d)×H^D1(Ω,d)({\overline{{\bf u}}}^{t},{\widehat{{\bf u}}}^{t})\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), where 𝐮¯t=𝒮12(φt){\overline{{\bf u}}}^{t}=\mathcal{S}_{1}^{2}(\varphi^{t}) and 𝐮^t=𝒮(φt){\widehat{{\bf u}}}^{t}=\mathcal{S}(\varphi^{t}).

Setting (φ0,𝐮¯0,𝐮^0)=(φ,𝐮¯,𝐮^)(\varphi_{0},{\overline{{\bf u}}}_{0},{\widehat{{\bf u}}}_{0})=(\varphi,{\overline{{\bf u}}},{\widehat{{\bf u}}}), by following a similar proof to [14, Lem. 25], we define for τ0>0\tau_{0}>0 sufficiently small the function 𝑭1:(τ0,τ0)×H¯D1(Ω,d)(H¯D1(Ω,d)){\bm{F}}_{1}:(-\tau_{0},\tau_{0})\times\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\to(\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*} by

𝑭1(t,𝐮)[𝐯]\displaystyle{\bm{F}}_{1}(t,{\bf u})[{\bf v}] =Ω(¯(φ)(Tt1𝐮)sym:(Tt1𝐯)sym((𝐅¯Tt)𝐯))det(Tt)dx\displaystyle=\int_{\Omega}\big{(}\overline{{\mathbb{C}}}(\varphi)(\nabla T_{t}^{-1}\nabla{\bf u})^{\rm sym}:(\nabla T_{t}^{-1}\nabla{\bf v})^{\rm sym}-((\overline{{\bf F}}\circ T_{t})\cdot{\bf v})\big{)}\det(\nabla T_{t})\,\mathrm{dx}
Γ¯N((𝐠¯Tt)𝐯det(Tt)Tt𝐧dd1.\displaystyle\quad-\int_{\overline{\Gamma}_{N}}((\overline{{\bf g}}\circ T_{t})\cdot{\bf v}\det(\nabla T_{t})\|\nabla T_{t}^{-\top}{\bf n}\|\,\mathrm{d}\mathcal{H}^{d-1}.

Using a change of variables y=Tt(𝒙)y=T_{t}(\bm{x}), the relation

Tt1(𝐮~Tt)=(𝐮~)Tt for 𝐮~:Tt(Ω)d,\nabla T_{t}^{-1}\nabla(\widetilde{{\bf u}}\circ T_{t})=(\nabla\widetilde{{\bf u}})\circ T_{t}\quad\text{ for }\widetilde{{\bf u}}:T_{t}(\Omega)\to{\mathbb{R}}^{d},

and also [57, Prop. 2.47] for the boundary transformation, we obtain

𝑭1(t,𝐮)[𝐯]\displaystyle{\bm{F}}_{1}(t,{\bf u})[{\bf v}] =Tt(Ω)¯(φTt1)y(𝐮Tt1):y(𝐯Tt1)dy\displaystyle=\int_{T_{t}(\Omega)}\overline{{\mathbb{C}}}(\varphi\circ T_{t}^{-1})\mathcal{E}_{y}({\bf u}\circ T_{t}^{-1}):\mathcal{E}_{y}({\bf v}\circ T_{t}^{-1})\,\mathrm{d}y
Tt(Ω)𝐅¯(𝐯Tt1)dyTt(Γ¯N)𝐠¯(𝐯Tt1)dyd1,\displaystyle\quad-\int_{T_{t}(\Omega)}\overline{{\bf F}}\cdot({\bf v}\circ T_{t}^{-1})\,\mathrm{d}y-\int_{T_{t}(\overline{\Gamma}_{N})}\overline{{\bf g}}\cdot({\bf v}\circ T_{t}^{-1})\,\mathrm{d}\mathcal{H}^{d-1}_{y},

where y(𝐮)=12(y𝐮+(y𝐮))\mathcal{E}_{y}({\bf u})=\frac{1}{2}(\nabla_{y}{\bf u}+(\nabla_{y}{\bf u})^{\top}) for 𝐮:Tt(Ω)d{\bf u}:T_{t}(\Omega)\to{\mathbb{R}}^{d} and dyd1\,\mathrm{d}\mathcal{H}^{d-1}_{y} denotes the (d1)(d-1)-dimensional Hausdorff measure related to 𝒚\bm{y}. From the properties of the mapping TtT_{t}, we find that Tt(Γ¯N)=Γ¯NT_{t}(\overline{\Gamma}_{N})=\overline{\Gamma}_{N} and 𝐯~:=𝐯Tt1H¯D1(Tt(Ω),d)\widetilde{{\bf v}}:={\bf v}\circ T_{t}^{-1}\in\overline{H}^{1}_{D}(T_{t}(\Omega),{\mathbb{R}}^{d}). Hence, from the above identity we observe that

𝑭1(t,𝐮¯tTt)[𝐯]=Ω¯(φt)y(𝐮¯t):y(𝐯~)𝐅¯𝐯~dyΓ¯N𝐠¯𝐯~dyd1=0{\bm{F}}_{1}(t,{\overline{{\bf u}}}^{t}\circ T_{t})[{\bf v}]=\int_{\Omega}\overline{{\mathbb{C}}}(\varphi^{t})\mathcal{E}_{y}({\overline{{\bf u}}}^{t}):\mathcal{E}_{y}(\widetilde{{\bf v}})-\overline{{\bf F}}\cdot\widetilde{{\bf v}}\,\mathrm{d}y-\int_{\overline{\Gamma}_{N}}\overline{{\bf g}}\cdot\widetilde{{\bf v}}\,\mathrm{d}\mathcal{H}^{d-1}_{y}=0

for all 𝐯H¯D1(Ω,d){\bf v}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Denoting by D𝐮𝑭1D_{\bf u}{\bm{F}}_{1} as the partial derivative of 𝑭1{\bm{F}}_{1} with respect to its second argument, we find that D𝐮𝑭1(0,𝐮):H¯D1(Ω,d)(H¯D1(Ω,d))D_{\bf u}{\bm{F}}_{1}(0,{\bf u}):\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\to(\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*} is given by

(D𝐮𝑭1(0,𝐮)[𝐯])[𝐰]=Ω¯(φ)(𝐯):(𝐰)dx\Big{(}D_{\bf u}{\bm{F}}_{1}(0,{\bf u})[{\bf v}]\Big{)}[{\bf w}]=\int_{\Omega}\overline{{\mathbb{C}}}(\varphi)\mathcal{E}({\bf v}):\mathcal{E}({\bf w})\,\mathrm{dx}

for all 𝐯,𝐰H¯D1(Ω,d){\bf v},{\bf w}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), where we have used the relation Tt1|t=0=𝕀\nabla T_{t}^{-1}|_{t=0}=\mathbb{I} the identity matrix. As D𝐮𝑭1(0,𝐮)D_{\bf u}{\bm{F}}_{1}(0,{\bf u}) is an isomorphism by the Lax–Milgram theorem, the application of the implicit function theorem allows us to infer that the mapping

t(𝐮¯tTt)t\mapsto({\overline{{\bf u}}}^{t}\circ T_{t})

is differentiable at t=0t=0 with derivative 𝐮¯˙[𝐕]:=t|t=0(𝐮¯tTt)H¯D1(Ω,d)\dot{{\overline{{\bf u}}}}[{\bf V}]:=\partial_{t}|_{t=0}({\overline{{\bf u}}}^{t}\circ T_{t})\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) being the unique solution to the distributional equation

D𝐮𝑭1(0,𝐮¯)[𝐮¯˙[𝐕]]=t𝑭1(0,𝐮¯) in (H¯D1(Ω,d)),D_{\bf u}{\bm{F}}_{1}(0,{\overline{{\bf u}}})\big{[}\dot{{\overline{{\bf u}}}}[{\bf V}]\big{]}=-\partial_{t}{\bm{F}}_{1}(0,{\overline{{\bf u}}})\text{ in }(\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*},

which reads as

Ω¯(φ)(𝐮¯˙[𝐕]):(𝜻)dx=Ω¯(φ)(𝐕(0)𝐮¯)sym:(𝜻)dx\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}(\varphi)\mathcal{E}(\dot{{\overline{{\bf u}}}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\int_{\Omega}\overline{{\mathbb{C}}}(\varphi)(\nabla{\bf V}(0)\nabla{\overline{{\bf u}}})^{\mathrm{sym}}:\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} (5.11)
+Ω¯(φ)(𝐮¯):(𝐕(0)𝜻)sym¯(φ)(𝐮¯):(𝜻)div𝐕(0)dx\displaystyle\quad+\int_{\Omega}\overline{{\mathbb{C}}}(\varphi)\mathcal{E}({\overline{{\bf u}}}):(\nabla{\bf V}(0)\nabla{\bm{\zeta}})^{\mathrm{sym}}-\overline{{\mathbb{C}}}(\varphi)\mathcal{E}({\overline{{\bf u}}}):\mathcal{E}({\bm{\zeta}})\mathop{\rm div}\nolimits{\bf V}(0)\,\mathrm{dx}
+Ω(𝐅¯𝐕(0)+𝐅¯div(𝐕(0)))𝜻dx\displaystyle\quad+\int_{\Omega}\Big{(}\nabla\overline{{\bf F}}{\bf V}(0)+\overline{{\bf F}}\mathop{\rm div}\nolimits({\bf V}(0))\Big{)}\cdot{\bm{\zeta}}\,\mathrm{dx}
+Γ¯N(𝐠¯𝐕(0)+𝐠¯(div(𝐕(0))𝐧𝐕(0)𝐧))𝜻dd1\displaystyle\quad+\int_{\overline{\Gamma}_{N}}\Big{(}\nabla\overline{{\bf g}}{\bf V}(0)+\overline{{\bf g}}\Big{(}\mathop{\rm div}\nolimits({\bf V}(0))-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\Big{)}\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}

for all 𝜻H¯D1(Ω,d){\bm{\zeta}}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). In the above, we have made use of the following relations (see [57, Lem. 2.31, Prop. 2.36, Lem. 2.49])

tTt|t=0=𝐕(0),tTt1|t=0=𝐕(0),\displaystyle\partial_{t}\nabla T_{t}|_{t=0}=\nabla{\bf V}(0),\quad\partial_{t}\nabla T_{t}^{-1}|_{t=0}=-\nabla{\bf V}(0),
tdetTt|t=0=div𝐕(0),t(fTt)|t=0=f𝐕(0),\displaystyle\partial_{t}\det\nabla T_{t}|_{t=0}=\mathop{\rm div}\nolimits{\bf V}(0),\quad\partial_{t}(f\circ T_{t})|_{t=0}=\nabla f\cdot{\bf V}(0),
t(det(Tt)(Tt)1𝐧)|t=0=div𝐕(0)𝐧𝐕(0)𝐧.\displaystyle\partial_{t}(\det(\nabla T_{t})\|(\nabla T_{t})^{-1}\cdot{\bf n}\|)|_{t=0}=\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}.

Furthermore, substituting 𝜻=𝐮¯˙[𝐕]{\bm{\zeta}}=\dot{{\overline{{\bf u}}}}[{\bf V}] into (5.11), by means of Korn’s inequality and the smoothness of 𝐕(0){\bf V}(0), we obtain the estimate

𝐮¯˙[𝐕]H1(Ω)C(𝐮¯H1(Ω)+𝐅¯H1(Ω)+𝐠¯H2(Γ¯N)).\displaystyle\mathopen{\|}\dot{{\overline{{\bf u}}}}[{\bf V}]\mathclose{\|}_{H^{1}(\Omega)}\leq C\Big{(}\|{\overline{{\bf u}}}\|_{H^{1}(\Omega)}+\|\overline{{\bf F}}\|_{H^{1}(\Omega)}+\|\overline{{\bf g}}\|_{H^{2}(\overline{\Gamma}_{N})}\Big{)}. (5.12)

Via a similar procedure, for a small τ0>0\tau_{0}>0, we consider 𝑭2:(τ0,τ0)×H^D1(Ω,d)(H^D1(Ω,d)){\bm{F}}_{2}:(-\tau_{0},\tau_{0})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\to(\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*} defined as

𝑭2(t,𝐮)[𝐯]\displaystyle{\bm{F}}_{2}(t,{\bf u})[{\bf v}] =Ω(^(φ)(Tt1𝐮)sym:(Tt1𝐯)sym((𝐅^Tt)𝐯))det(Tt)dx\displaystyle=\int_{\Omega}\big{(}\widehat{{\mathbb{C}}}(\varphi)(\nabla T_{t}^{-1}\nabla{\bf u})^{\rm sym}:(\nabla T_{t}^{-1}\nabla{\bf v})^{\rm sym}-((\widehat{{\bf F}}\circ T_{t})\cdot{\bf v})\big{)}\det(\nabla T_{t})\,\mathrm{dx}
Γ^N((𝐠^Tt)𝐯det(Tt)Tt𝐧dd1\displaystyle\quad-\int_{\widehat{\Gamma}_{N}}((\widehat{{\bf g}}\circ T_{t})\cdot{\bf v}\det(\nabla T_{t})\|\nabla T_{t}^{-\top}{\bf n}\|\,\mathrm{d}\mathcal{H}^{d-1}
Ωχ(φ)^(φ)(Tt1(𝐮¯tTt))sym:(Tt1𝐯)symdet(Tt)dx.\displaystyle\quad-\int_{\Omega}\chi(\varphi)\widehat{{\mathbb{C}}}(\varphi)(\nabla T_{t}^{-1}\nabla({\overline{{\bf u}}}^{t}\circ T_{t}))^{\rm sym}:(\nabla T_{t}^{-1}\nabla{\bf v})^{\rm sym}\det(\nabla T_{t})\,\mathrm{dx}.

Then, by a change of variables 𝒚=Tt(𝒙)\bm{y}=T_{t}(\bm{x}), we find that

𝑭2(t,𝐮^tTt)[𝐯]\displaystyle{\bm{F}}_{2}(t,{\widehat{{\bf u}}}^{t}\circ T_{t})[{\bf v}] =Ω^(φt)y(𝐮^t):y(𝐯~)𝐅^𝐯~dyΓ^N𝐠^𝐯~dyd1\displaystyle=\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi^{t})\mathcal{E}_{y}({\widehat{{\bf u}}}^{t}):\mathcal{E}_{y}(\widetilde{{\bf v}})-\widehat{{\bf F}}\cdot\widetilde{{\bf v}}\,\mathrm{d}y-\int_{\widehat{\Gamma}_{N}}\widehat{{\bf g}}\cdot\widetilde{{\bf v}}\,\mathrm{d}\mathcal{H}^{d-1}_{y}
Ωχ(φt)^(φt)y(𝐮¯t):y(𝐯~)dy=0\displaystyle\quad-\int_{\Omega}\chi(\varphi^{t})\widehat{{\mathbb{C}}}(\varphi^{t})\mathcal{E}_{y}({\overline{{\bf u}}}^{t}):\mathcal{E}_{y}(\widetilde{{\bf v}})\,\mathrm{d}y=0

where 𝐯~=𝐯Tt1H^D1(Tt(Ω),d)\widetilde{{\bf v}}={\bf v}\circ T_{t}^{-1}\in\widehat{H}^{1}_{D}(T_{t}(\Omega),{\mathbb{R}}^{d}). Since D𝐮𝑭2(0,𝐮):H^D1(Ω,d)(H^D1(Ω,d))D_{\bf u}{\bm{F}}_{2}(0,{\bf u}):\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\to(\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*} given by

(D𝐮𝑭2(0,𝐮)[𝐯])[𝐰]=Ω^(φ)(𝐯):(𝐰)dx for all 𝐯,𝐰H^D1(Ω,d)\Big{(}D_{\bf u}{\bm{F}}_{2}(0,{\bf u})[{\bf v}]\Big{)}[{\bf w}]=\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi)\mathcal{E}({\bf v}):\mathcal{E}({\bf w})\,\mathrm{dx}\quad\text{ for all }{\bf v},{\bf w}\in\widehat{H}_{D}^{1}(\Omega,{\mathbb{R}}^{d})

is an isomorphism, by the implicit function theorem we infer that the mapping

t(𝐮^tTt)t\mapsto({\widehat{{\bf u}}}^{t}\circ T_{t})

is differentiable at t=0t=0 with derivative 𝐮^˙[𝐕]:=t|t=0(𝐮^tTt)H^D1(Ω,d)\dot{{\widehat{{\bf u}}}}[{\bf V}]:=\partial_{t}|_{t=0}({\widehat{{\bf u}}}^{t}\circ T_{t})\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) being the unique solution to the distributional equation

D𝐮𝑭2(0,𝐮^)[𝐮^˙[𝐕]]=t𝑭2(0,𝐮^) in (H^D1(Ω,d)),D_{\bf u}{\bm{F}}_{2}(0,{\widehat{{\bf u}}})\big{[}\dot{{\widehat{{\bf u}}}}[{\bf V}]\big{]}=-\partial_{t}{\bm{F}}_{2}(0,{\widehat{{\bf u}}})\text{ in }(\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}))^{*},

which reads as

Ω^(φ)(𝐮^˙[𝐕]):(𝜻)dx=Ω^(φ)(𝐕(0)𝐮^χ(φ)𝐕(0)𝐮¯)sym:(𝜻)dx\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi)\mathcal{E}(\dot{{\widehat{{\bf u}}}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi)(\nabla{\bf V}(0)\nabla{\widehat{{\bf u}}}-\chi(\varphi)\nabla{\bf V}(0)\nabla{\overline{{\bf u}}})^{\mathrm{sym}}:\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} (5.13)
+Ω^(φ)((𝐮^)χ(φ)(𝐮¯)):(𝐕(0)𝜻)sym+χ(φ)^(φ)(𝐮¯˙[𝐕]):(𝜻)dx\displaystyle\quad+\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi)(\mathcal{E}({\widehat{{\bf u}}})-\chi(\varphi)\mathcal{E}({\overline{{\bf u}}})):(\nabla{\bf V}(0)\nabla{\bm{\zeta}})^{\mathrm{sym}}+\chi(\varphi)\widehat{{\mathbb{C}}}(\varphi)\mathcal{E}(\dot{{\overline{{\bf u}}}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}
Ω^(φ)((𝐮^)χ(φ)(𝐮¯)):(𝜻)div𝐕(0)dx\displaystyle\quad-\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi)(\mathcal{E}({\widehat{{\bf u}}})-\chi(\varphi)\mathcal{E}({\overline{{\bf u}}})):\mathcal{E}({\bm{\zeta}})\mathop{\rm div}\nolimits{\bf V}(0)\,\mathrm{dx}
+Ω(𝐅^𝐕(0)+𝐅^div(𝐕(0)))𝜻dx\displaystyle\quad+\int_{\Omega}\Big{(}\nabla\widehat{{\bf F}}{\bf V}(0)+\widehat{{\bf F}}\mathop{\rm div}\nolimits({\bf V}(0))\Big{)}\cdot{\bm{\zeta}}\,\mathrm{dx}
+Γ^N(𝐠^𝐕(0)+𝐠^(div(𝐕(0))𝐧𝐕(0)𝐧))𝜻dd1\displaystyle\quad+\int_{\widehat{\Gamma}_{N}}\Big{(}\nabla\widehat{{\bf g}}{\bf V}(0)+\widehat{{\bf g}}\Big{(}\mathop{\rm div}\nolimits({\bf V}(0))-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\Big{)}\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}

for all 𝜻H^D1(Ω,d){\bm{\zeta}}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}). Substituting 𝜻=𝐮^˙[𝐕]{\bm{\zeta}}=\dot{{\widehat{{\bf u}}}}[{\bf V}] into (5.13), then using (5.12) and Korn’s inequality, we obtain the estimate

𝐮^˙[𝐕]H1(Ω)\displaystyle\mathopen{\|}\dot{{\widehat{{\bf u}}}}[{\bf V}]\mathclose{\|}_{H^{1}(\Omega)} C(𝐮^H1(Ω)+𝐮¯H1(Ω))\displaystyle\leq C\Big{(}\|{\widehat{{\bf u}}}\|_{H^{1}(\Omega)}+\|{\overline{{\bf u}}}\|_{H^{1}(\Omega)}\Big{)} (5.14)
+C(𝐅¯H1(Ω)+𝐠¯H2(Γ¯N)+𝐅^H1(Ω)+𝐠^H2(Γ^N)).\displaystyle\quad+C\Big{(}\|\overline{{\bf F}}\|_{H^{1}(\Omega)}+\|\overline{{\bf g}}\|_{H^{2}(\overline{\Gamma}_{N})}+\|\widehat{{\bf F}}\|_{H^{1}(\Omega)}+\|\widehat{{\bf g}}\|_{H^{2}(\widehat{\Gamma}_{N})}\Big{)}.

The next result details an optimality condition for minimisers φε\varphi^{\varepsilon} to (𝑷ε)(\bm{P}^{\varepsilon}) obtained via geometric variations.

Theorem 5.1.

Assume (A1)(A6), and additionally suppose that 𝐅¯,𝐅^H1(Ω,d)\overline{{\bf F}},\widehat{{\bf F}}\in H^{1}(\Omega,{\mathbb{R}}^{d}), 𝐔¯=𝐔^=𝟎,𝐠¯H2(Γ¯N,d)\overline{\bm{U}}=\widehat{\bm{U}}=\bm{0},\overline{{\bf g}}\in H^{2}(\overline{\Gamma}_{N},{\mathbb{R}}^{d}), 𝐠^H2(Γ^N,d)\widehat{{\bf g}}\in H^{2}(\widehat{\Gamma}_{N},{\mathbb{R}}^{d}) and 𝐮tarH2(Γtar,d){\bf u}^{\rm tar}\in H^{2}(\Gamma^{\rm tar},{\mathbb{R}}^{d}) hold. Let φε𝒰ad\varphi^{\varepsilon}\in\mathcal{U}_{\rm ad} be a minimiser to (𝐏ε)(\bm{P}^{\varepsilon}), with corresponding solutions 𝐮¯ε=𝒮12(φε){\overline{{\bf u}}}^{\varepsilon}=\mathcal{S}_{1}^{2}(\varphi^{\varepsilon}) and 𝐮^ε=𝒮(φε){\widehat{{\bf u}}}^{\varepsilon}=\mathcal{S}(\varphi^{\varepsilon}). For every admissible velocity 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}}, let (𝐮¯ε˙[𝐕],𝐮^ε˙[𝐕])H¯D1(Ω,d)×H^D1(Ω,d)(\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}],\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}])\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d})\times\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) denote the unique solutions to (5.11) and (5.13) with (φ,𝐮¯,𝐮^)=(φε,𝐮¯ε,𝐮^ε)(\varphi,{\overline{{\bf u}}},{\widehat{{\bf u}}})=(\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon}). Then, for all 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}}, it holds that

0\displaystyle 0 =12ΓtarW(𝐮^ε𝐮tar)(𝐮^ε𝐮tar)(div𝐕(0)𝐧𝐕(0)𝐧)dd1\displaystyle=\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\,\mathrm{d}\mathcal{H}^{d-1} (5.15)
+ΓtarW(𝐮^ε𝐮tar)(𝐮^ε˙[𝐕](𝐮tar)𝐕(0))dd1\displaystyle\quad+\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot(\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]-(\nabla{\bf u}^{\rm tar}){\bf V}(0))\,\mathrm{d}\mathcal{H}^{d-1}
+γΩ(ε2|φε|2+1εΨ(φε))div𝐕(0)εφε𝐕(0)φεdx.\displaystyle\quad+\gamma\int_{\Omega}\Big{(}\frac{\varepsilon}{2}|\nabla\varphi^{\varepsilon}|^{2}+\frac{1}{\varepsilon}\Psi(\varphi^{\varepsilon})\Big{)}\mathop{\rm div}\nolimits{\bf V}(0)-\varepsilon\nabla\varphi^{\varepsilon}\cdot\nabla{\bf V}(0)\nabla\varphi^{\varepsilon}\,\mathrm{dx}.
Remark 5.2.

With more regularity, it is possible to relate (5.15) to the optimality condition (4.21), see [14, 41] for more details.

Proof.

For any 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}}, let T𝒯adT\in\mathcal{T}_{\mathrm{ad}} be the associated transformation and consider the scalar function g(t):=Jredε(φεTt1)g(t):=J_{\rm red}^{\varepsilon}(\varphi^{\varepsilon}\circ T_{t}^{-1}) for t(τ0,τ0)t\in(-\tau_{0},\tau_{0}), where τ0\tau_{0} is sufficiently small. As φε\varphi^{\varepsilon} is a minimiser of JredεJ_{\rm red}^{\varepsilon}, we have

g(0)=ddtJredε(φεTt1)|t=0=0.g^{\prime}(0)=\frac{d}{dt}J_{\rm red}^{\varepsilon}(\varphi^{\varepsilon}\circ T_{t}^{-1})|_{t=0}=0.

The directional derivative

ddtEε(φεTt1)|t=0=Ω(ε2|φε|2+1εΨ(φε))div𝐕(0)εφε𝐕(0)φεdx\frac{d}{dt}E_{\varepsilon}(\varphi^{\varepsilon}\circ T_{t}^{-1})|_{t=0}=\int_{\Omega}\Big{(}\frac{\varepsilon}{2}|\nabla\varphi^{\varepsilon}|^{2}+\frac{1}{\varepsilon}\Psi(\varphi^{\varepsilon})\Big{)}\mathop{\rm div}\nolimits{\bf V}(0)-\varepsilon\nabla\varphi^{\varepsilon}\cdot\nabla{\bf V}(0)\nabla\varphi^{\varepsilon}\,\mathrm{dx}

can be obtained as in [41, Lem. 7.5]. Denoting by 𝐮^ε(t)=𝒮(φεTt1){\widehat{{\bf u}}}^{\varepsilon}(t)=\mathcal{S}(\varphi^{\varepsilon}\circ T_{t}^{-1}), then the derivative of G(φεTt1)G(\varphi^{\varepsilon}\circ T_{t}^{-1}) can be obtained by a standard change of variables:

ddtG(φεTt1)\displaystyle\frac{d}{dt}G(\varphi^{\varepsilon}\circ T_{t}^{-1})
=12ddtΓtarW((𝐮^ε(t)𝐮tar)Tt)((𝐮^ε(t)𝐮tar)Tt)det(Tt)Tt𝐧dd1\displaystyle\quad=\frac{1}{2}\frac{d}{dt}\int_{\Gamma^{\rm tar}}W(({\widehat{{\bf u}}}^{\varepsilon}(t)-{\bf u}^{\rm tar})\circ T_{t})\cdot(({\widehat{{\bf u}}}^{\varepsilon}(t)-{\bf u}^{\rm tar})\circ T_{t})\det(\nabla T_{t})\|\nabla T_{t}^{-\top}{\bf n}\|\,\mathrm{d}\mathcal{H}^{d-1}
=ΓtarW(𝐮^ε𝐮tar)ddt(𝐮^ε(t)Tt𝐮tarTt)|t=0dd1\displaystyle\quad=\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot\frac{d}{dt}({\widehat{{\bf u}}}^{\varepsilon}(t)\circ T_{t}-{\bf u}^{\rm tar}\circ T_{t})|_{t=0}\,\mathrm{d}\mathcal{H}^{d-1}
+12ΓtarW(𝐮^ε𝐮tar)(𝐮^ε𝐮tar)ddt(det(Tt)Tt𝐧)dd1\displaystyle\qquad+\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\frac{d}{dt}(\det(\nabla T_{t})\|\nabla T_{t}^{-\top}{\bf n}\|)\,\mathrm{d}\mathcal{H}^{d-1}
=ΓtarW(𝐮^ε𝐮tar)(𝐮^ε˙[𝐕](𝐮tar)𝐕(0))dd1\displaystyle\quad=\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot(\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]-(\nabla{\bf u}^{\rm tar}){\bf V}(0))\,\mathrm{d}\mathcal{H}^{d-1}
+12ΓtarW(𝐮^ε𝐮tar)(𝐮^ε𝐮tar)(div𝐕(0)𝐧𝐕(0)𝐧)dd1\displaystyle\qquad+\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\,\mathrm{d}\mathcal{H}^{d-1}

leading to (5.15). ∎

The convergence of (5.15) to the sharp interface limit ε0\varepsilon\to 0 is formulated as follows.

Theorem 5.2.

Suppose the hypotheses of Theorem 5.1 hold, and let φε𝒰ad\varphi^{\varepsilon}\in\mathcal{U}_{\rm ad} be a minimiser to (𝐏ε)(\bf P^{\varepsilon}). For any 𝐕𝒱ad{\bf V}\in\mathcal{V}_{\mathrm{ad}} with corresponding transformation T𝒯adT\in\mathcal{T}_{\mathrm{ad}}, there exists a non-relabelled subsequence ε0\varepsilon\to 0 such that

φε\displaystyle\varphi^{\varepsilon} φ0 in L1(Ω),Jredε(φε)Jred0(φ0) in ,\displaystyle\to\varphi_{0}\text{ in }L^{1}(\Omega),\quad J_{\rm red}^{\varepsilon}(\varphi^{\varepsilon})\to J_{\rm red}^{0}(\varphi_{0})\text{ in }{\mathbb{R}},
𝐮¯ε˙[𝐕]\displaystyle\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}] 𝐮¯˙0[𝐕] in H¯D1(Ω,d),𝐮^ε˙[𝐕]𝐮^˙0[𝐕] in H^D1(Ω,d),\displaystyle\rightharpoonup\dot{{\overline{{\bf u}}}}_{0}[{\bf V}]\text{ in }\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),\quad\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]\rightharpoonup\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}]\text{ in }\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}),

where φ0BV(Ω,{1,1})\varphi_{0}\in\mathrm{BV}(\Omega,\{-1,1\}) is a minimiser to the reduced functional Jred0(φ)=G(φ)+γTV(φ)J_{\rm red}^{0}(\varphi)=G(\varphi)+\gamma TV(\varphi), where the total variation TV(φ)TV(\varphi) for φBV(Ω)\varphi\in\mathrm{BV}(\Omega) is defined as

TV(φ)=sup{Ωφdivϕdx s.t. ϕC01(Ω,d),ϕL(Ω)1}.\displaystyle TV(\varphi)=\sup\Big{\{}\int_{\Omega}\varphi\mathop{\rm div}\nolimits\bm{\phi}\,\mathrm{dx}\text{ s.t. }\bm{\phi}\in C^{1}_{0}(\Omega,{\mathbb{R}}^{d}),\,\|\bm{\phi}\|_{L^{\infty}(\Omega)}\leq 1\Big{\}}.

Furthermore, 𝐮¯˙0[𝐕]\dot{{\overline{{\bf u}}}}_{0}[{\bf V}] and 𝐮^˙0[𝐕]\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}] satisfy (5.11) and (5.13), respectively, with (φ,𝐮¯,𝐮^)(\varphi,{\overline{{\bf u}}},{\widehat{{\bf u}}}) replaced by (φ0,𝐮¯0=𝒮12(φ0),𝐮^0=𝒮2(φ0))(\varphi_{0},{\overline{{\bf u}}}_{0}=\mathcal{S}_{1}^{2}(\varphi_{0}),{\widehat{{\bf u}}}_{0}=\mathcal{S}_{2}(\varphi_{0})). Lastly, it holds that

ddtJredε(φεTt1)|t=0ddtJred0(φ0Tt1)|t=0 in ,\displaystyle\frac{d}{dt}J_{\rm red}^{\varepsilon}(\varphi^{\varepsilon}\circ T_{t}^{-1})|_{t=0}\to\frac{d}{dt}J_{\rm red}^{0}(\varphi_{0}\circ T_{t}^{-1})|_{t=0}\text{ in }{\mathbb{R}}, (5.16)

where

ddtJred0(φ0Tt1)|t=0\displaystyle\frac{d}{dt}J_{\rm red}^{0}(\varphi_{0}\circ T_{t}^{-1})|_{t=0} (5.17)
=ΓtarW(𝐮^0𝐮tar)(𝐮^˙0[𝐕](𝐮tar)𝐕(0))dd1\displaystyle\quad=\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\cdot(\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}]-(\nabla{\bf u}^{\rm tar}){\bf V}(0))\,\mathrm{d}\mathcal{H}^{d-1}
+12ΓtarW(𝐮^0𝐮tar)(𝐮^0𝐮tar)(div𝐕(0)𝐧𝐕(0)𝐧)dd1\displaystyle\qquad+\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\,\mathrm{d}\mathcal{H}^{d-1}
+γΩ(div𝐕(0)μ𝐕(0)μ)d|D𝒳{φ0=1}|,\displaystyle\qquad+\gamma\int_{\Omega}\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-\mu\cdot\nabla{\bf V}(0)\mu\Big{)}d|\mathrm{D}\mathrm{\mathcal{X}}_{\{\varphi_{0}=1\}}|,

with μ=D𝒳{φ0=1}|D𝒳{φ0=1}|\mu=\frac{\mathrm{D}\mathrm{\mathcal{X}}_{\{\varphi_{0}=1\}}}{|\mathrm{D}\mathrm{\mathcal{X}}_{\{\varphi_{0}=1\}}|} as the generalised unit normal on the set {φ0=1}\{\varphi_{0}=1\}.

Remark 5.3.

With more regularity, it is possible to relate (5.17) with the solvability condition (5.9) in the two-phase setting, see [14, 41] for more details.

Proof.

The first two assertions on the convergence of φε\varphi^{\varepsilon} and Jredε(φε)J_{\rm red}^{\varepsilon}(\varphi^{\varepsilon}) come from Lemma 5.1. Consequently, by the calculations in the proof of [27, Thm. 4.2] we infer the convergence, as ε0\varepsilon\to 0,

Ω(ε2|φε|2+1εΨ(φε))div𝐕(0)εφε𝐕(0)φεdx\displaystyle\int_{\Omega}\Big{(}\frac{\varepsilon}{2}|\nabla\varphi^{\varepsilon}|^{2}+\frac{1}{\varepsilon}\Psi(\varphi^{\varepsilon})\Big{)}\mathop{\rm div}\nolimits{\bf V}(0)-\varepsilon\nabla\varphi^{\varepsilon}\cdot\nabla{\bf V}(0)\nabla\varphi^{\varepsilon}\,\mathrm{dx}
Ω(div𝐕(0)μ𝐕(0)μ)d|D𝒳{φ0=1}|.\displaystyle\quad\to\int_{\Omega}\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-\mu\cdot\nabla{\bf V}(0)\mu\Big{)}d|\mathrm{D}\mathrm{\mathcal{X}}_{\{\varphi_{0}=1\}}|.

Next, from (5.11) and (5.13), we see that 𝐮¯ε˙[𝐕]\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}] and 𝐮^ε˙[𝐕]\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}] satisfy

Ω¯(φε)(𝐮¯ε˙[𝐕]):(𝜻)dx=φε,𝐮¯ε(𝜻),Ω^(φε)(𝐮^ε˙[𝐕]):(𝜻)dx=φε,𝐮¯ε,𝐮^ε(𝜻),\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}(\varphi^{\varepsilon})\mathcal{E}(\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon}}({\bm{\zeta}}),\quad\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi^{\varepsilon})\mathcal{E}(\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}=\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon}}({\bm{\zeta}}),

where φε,𝐮¯ε(𝜻)\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon}}({\bm{\zeta}}) and φε,𝐮¯ε,𝐮^ε(𝜻)\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon}}({\bm{\zeta}}) denotes the right-hand sides of (5.11) and (5.13), respectively. Thanks to Corollary 3.1, along a non-relabelled subsequence, 𝐮¯ε𝐮¯0H¯D1(Ω,d){\overline{{\bf u}}}^{\varepsilon}\to{\overline{{\bf u}}}_{0}\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝐮^ε𝐮^0H^D1(Ω,d){\widehat{{\bf u}}}^{\varepsilon}\to{\widehat{{\bf u}}}_{0}\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) strongly as ε0\varepsilon\to 0. Hence, together with the dominated convergence theorem, it is clear that, as ε0\varepsilon\to 0,

φε,𝐮¯ε(𝜻)φ0,𝐮¯0(𝜻),φε,𝐮¯ε,𝐮^ε(𝜻)φ0,𝐮¯0,𝐮^0(𝜻).\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon}}({\bm{\zeta}})\to\mathcal{F}_{\varphi_{0},{\overline{{\bf u}}}_{0}}({\bm{\zeta}}),\quad\mathcal{F}_{\varphi^{\varepsilon},{\overline{{\bf u}}}^{\varepsilon},{\widehat{{\bf u}}}^{\varepsilon}}({\bm{\zeta}})\to\mathcal{F}_{\varphi_{0},{\overline{{\bf u}}}_{0},{\widehat{{\bf u}}}_{0}}({\bm{\zeta}}).

On the other hand, we infer from (5.12) and (5.14) that 𝐮¯ε˙[𝐕]\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}] and 𝐮^ε˙[𝐕]\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}] are uniformly bounded in H¯D1(Ω,d)\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and H^D1(Ω,d)\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), which then implies the existence of limit functions 𝐮¯˙0[𝐕]H¯D1(Ω,d)\dot{{\overline{{\bf u}}}}_{0}[{\bf V}]\in\overline{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}) and 𝐮^˙0[𝐕]H^D1(Ω,d)\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}]\in\widehat{H}^{1}_{D}(\Omega,{\mathbb{R}}^{d}), where

Ω¯(φε)(𝐮¯ε˙[𝐕]):(𝜻)dx\displaystyle\int_{\Omega}\overline{{\mathbb{C}}}(\varphi^{\varepsilon})\mathcal{E}(\dot{{\overline{{\bf u}}}^{\varepsilon}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} Ω¯(φ0)(𝐮¯˙0[𝐕]):(𝜻)dx,\displaystyle\to\int_{\Omega}\overline{{\mathbb{C}}}(\varphi_{0})\mathcal{E}(\dot{{\overline{{\bf u}}}}_{0}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx},
Ω^(φε)(𝐮^ε˙[𝐕]):(𝜻)dx\displaystyle\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi^{\varepsilon})\mathcal{E}(\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx} Ω^(φ0)(𝐮^˙0[𝐕]):(𝜻)dx.\displaystyle\to\int_{\Omega}\widehat{{\mathbb{C}}}(\varphi_{0})\mathcal{E}(\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}]):\mathcal{E}({\bm{\zeta}})\,\mathrm{dx}.

Lastly, using the compactness of the boundary-trace operator, we obtain, as ε0\varepsilon\to 0,

ΓtarW(𝐮^ε𝐮tar)(𝐮^ε˙[𝐕](𝐮tar)𝐕(0))dd1\displaystyle\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot(\dot{{\widehat{{\bf u}}}^{\varepsilon}}[{\bf V}]-(\nabla{\bf u}^{\rm tar}){\bf V}(0))\,\mathrm{d}\mathcal{H}^{d-1}
+12ΓtarW(𝐮^ε𝐮tar)(𝐮^ε𝐮tar)(div𝐕(0)𝐧𝐕(0)𝐧)dd1\displaystyle\qquad+\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}^{\varepsilon}-{\bf u}^{\rm tar})\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\,\mathrm{d}\mathcal{H}^{d-1}
ΓtarW(𝐮^0𝐮tar)(𝐮^˙0[𝐕](𝐮tar)𝐕(0))dd1\displaystyle\quad\to\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\cdot(\dot{{\widehat{{\bf u}}}}_{0}[{\bf V}]-(\nabla{\bf u}^{\rm tar}){\bf V}(0))\,\mathrm{d}\mathcal{H}^{d-1}
+12ΓtarW(𝐮^0𝐮tar)(𝐮^0𝐮tar)(div𝐕(0)𝐧𝐕(0)𝐧)dd1.\displaystyle\qquad+\frac{1}{2}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\cdot({\widehat{{\bf u}}}_{0}-{\bf u}^{\rm tar})\Big{(}\mathop{\rm div}\nolimits{\bf V}(0)-{\bf n}\cdot\nabla{\bf V}(0){\bf n}\Big{)}\,\mathrm{d}\mathcal{H}^{d-1}.

This shows (5.16) and completes the proof. ∎

6 Numerical simulations

In this section we present the finite element discretisation and showcase several numerical simulations in two and three dimensions for the two-phase case. Namely, we have L=2L=2 and we consider the optimal distribution of a single type of active material within a passive material.

6.1 Finite element discretisation

We assume that Ω\Omega is a polyhedral domain and let 𝒯h\mathcal{T}_{h} be a regular triangulation of Ω\Omega into disjoint open simplices. Associated with 𝒯h\mathcal{T}_{h} are the piecewise linear finite element spaces

Sh={ζC0(Ω¯):ζ|oP1(o)o𝒯h}andSSh=Sh××Sh=[Sh]d,\displaystyle S^{h}=\left\{\zeta\in C^{0}(\overline{\Omega}):\,\zeta_{|_{o}}\in P_{1}(o)\,\forall o\in\mathcal{T}_{h}\right\}\quad\text{and}\quad\SS^{h}=S^{h}\times\cdots\times S^{h}=[S^{h}]^{d},

where we denote by P1(o)P_{1}(o) the set of all affine linear functions on oo, cf. [22]. In addition we define

𝒱h={ζSh:|ζ|1 in Ω¯},\mathcal{V}^{h}=\left\{\zeta\in S^{h}:|\zeta|\leq 1\text{ in }\overline{\Omega}\right\}, (6.1)

as well as

SS¯Dh={𝜼SSh:𝜼=𝟎 on Γ¯D},SS^Dh={𝜼SSh:𝜼=𝟎 on Γ^D}.\overline{\SS}^{h}_{D}=\left\{\bm{\eta}\in\SS^{h}:\bm{\eta}={\bf 0}\text{ on }\overline{\Gamma}_{D}\right\},\quad\widehat{\SS}^{h}_{D}=\left\{\bm{\eta}\in\SS^{h}:\bm{\eta}={\bf 0}\text{ on }\widehat{\Gamma}_{D}\right\}.

We also let (,)(\cdot,\cdot) denote the L2L^{2}–inner product on Ω\Omega, and let (,)h(\cdot,\cdot)^{h} be the usual mass lumped L2L^{2}–inner product on Ω\Omega associated with 𝒯h\mathcal{T}_{h}. In addition, 𝐀,𝐁=(𝐀,𝐁)\langle{\bf A},{\bf B}\rangle_{{\mathbb{C}}}=({\mathbb{C}}{\bf A},{\bf B}) for any fourth order tensor {\mathbb{C}} and any matrices 𝐀{\bf A} and 𝐁{\bf B}. Finally, τ\tau denotes a chosen uniform time step size.

We now introduce finite element approximations of the state equations (3.9) and (3.10), adjoint systems (4.18) and (4.20), as well as the optimality conditions (4.21) with a pseudo-time evolution based on an L2L^{2}-gradient flow approach. In particular, we consider the obstacle potential (5.10) with Ψ~(s)=12(1s2)\widetilde{\Psi}(s)=\frac{1}{2}(1-s^{2}), which leads to a variational inequality.

The fully discrete numerical scheme is formulated as follows: Given φhn1𝒱h\varphi_{h}^{n-1}\in\mathcal{V}^{h}, find (𝐮¯hn,𝐮^hn,𝐪^hn,𝒑¯hn,φhn)SS¯Dh×SS^Dh×SS^Dh×SS¯Dh×𝒱h({\overline{{\bf u}}}_{h}^{n},{\widehat{{\bf u}}}_{h}^{n},\widehat{{\bf q}}_{h}^{n},\overline{\bm{p}}_{h}^{n},\varphi_{h}^{n})\in\overline{\SS}^{h}_{D}\times\widehat{\SS}^{h}_{D}\times\widehat{\SS}^{h}_{D}\times\overline{\SS}^{h}_{D}\times\mathcal{V}^{h} such that

(𝐮¯hn),(𝜻)¯(φhn1)=(𝐅¯,𝜻)h+Γ¯N𝐠¯𝜻dd1𝜻SS¯Dh,\displaystyle\langle\mathcal{E}({\overline{{\bf u}}}_{h}^{n}),\mathcal{E}({\bm{\zeta}})\rangle_{\overline{{\mathbb{C}}}(\varphi_{h}^{n-1})}=\left(\overline{{\bf F}},{\bm{\zeta}}\right)^{h}+\int_{\overline{\Gamma}_{N}}\overline{{\bf g}}\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}\qquad\forall{\bm{\zeta}}\in\overline{\SS}^{h}_{D}, (6.2a)
(𝐮^hn)χ(φhn1)(𝐮¯hn),(𝜻)^(φhn1)=(𝐅^,𝜻)h+Γ^N𝐠^𝜻dd1𝜻SS^Dh,\displaystyle\langle\mathcal{E}({\widehat{{\bf u}}}_{h}^{n})-\chi(\varphi_{h}^{n-1})\mathcal{E}({\overline{{\bf u}}}_{h}^{n}),\mathcal{E}({\bm{\zeta}})\rangle_{\widehat{{\mathbb{C}}}(\varphi_{h}^{n-1})}=\big{(}\widehat{{\bf F}},{\bm{\zeta}}\big{)}^{h}+\int_{\widehat{\Gamma}_{N}}\widehat{{\bf g}}\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}\qquad\forall{\bm{\zeta}}\in\widehat{\SS}^{h}_{D}, (6.2b)
(𝐪^hn),(𝜻)^(φhn1)=ΓtarW(𝐮^hn𝐮tar)𝜻dd1𝜻SS^Dh,\displaystyle\langle\mathcal{E}(\widehat{{\bf q}}_{h}^{n}),\mathcal{E}({\bm{\zeta}})\rangle_{\widehat{{\mathbb{C}}}(\varphi_{h}^{n-1})}=\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{h}^{n}-{\bf u}^{\rm tar})\cdot{\bm{\zeta}}\,\mathrm{d}\mathcal{H}^{d-1}\qquad\forall{\bm{\zeta}}\in\widehat{\SS}^{h}_{D}, (6.2c)
(𝒑¯hn),(𝜻)¯(φhn1)=χ(φhn1)(𝐪^hn),(𝜻)^(φhn1)𝜻SS¯Dh,\displaystyle\langle\mathcal{E}(\overline{\bm{p}}_{h}^{n}),\mathcal{E}({\bm{\zeta}})\rangle_{\overline{{\mathbb{C}}}(\varphi_{h}^{n-1})}=\langle\chi(\varphi_{h}^{n-1})\mathcal{E}(\widehat{{\bf q}}_{h}^{n}),\mathcal{E}({\bm{\zeta}})\rangle_{\widehat{{\mathbb{C}}}(\varphi_{h}^{n-1})}\qquad\forall{\bm{\zeta}}\in\overline{\SS}^{h}_{D}, (6.2d)
(ετ(φhnφhn1)γεφhn,ζφhn)h+γε(φhn,(ζφhn))\displaystyle\left(\tfrac{\varepsilon}{\tau}(\varphi_{h}^{n}-\varphi_{h}^{n-1})-\tfrac{\gamma}{\varepsilon}\varphi_{h}^{n},\zeta-\varphi_{h}^{n}\right)^{h}+\gamma\varepsilon(\nabla\varphi_{h}^{n},\nabla(\zeta-\varphi_{h}^{n}))
+χ(φhn1)(𝐮¯hn),(ζφhn)(𝐪^hn)^(φhn1)\displaystyle\qquad\quad+\langle\chi^{\prime}(\varphi_{h}^{n-1})\mathcal{E}({\overline{{\bf u}}}_{h}^{n}),(\zeta-\varphi_{h}^{n})\mathcal{E}(\widehat{{\bf q}}_{h}^{n})\rangle_{\widehat{{\mathbb{C}}}(\varphi_{h}^{n-1})}
(𝐮^hn)χ(φhn1)(𝐮¯hn),(ζφhn)(𝐪^hn)^(φhn1)\displaystyle\qquad\quad-\langle\mathcal{E}({\widehat{{\bf u}}}_{h}^{n})-\chi(\varphi_{h}^{n-1})\mathcal{E}({\overline{{\bf u}}}_{h}^{n}),(\zeta-\varphi_{h}^{n})\mathcal{E}(\widehat{{\bf q}}_{h}^{n})\rangle_{\widehat{{\mathbb{C}}}^{\prime}(\varphi_{h}^{n-1})}
(𝐮¯hn),(ζφhn)(𝒑¯hn)¯(φhn1)0ζ𝒱h.\displaystyle\quad\qquad-\langle\mathcal{E}({\overline{{\bf u}}}_{h}^{n}),(\zeta-\varphi_{h}^{n})\mathcal{E}(\overline{\bm{p}}_{h}^{n})\rangle_{\overline{{\mathbb{C}}}^{\prime}(\varphi_{h}^{n-1})}\geq 0\qquad\forall\zeta\in\mathcal{V}^{h}. (6.2e)

We implemented the scheme (6.2) with the help of the finite element toolbox ALBERTA, see [56]. To increase computational efficiency, we employ adaptive meshes, which have a finer mesh size hfh_{f} within the diffuse interfacial regions and a coarser mesh size hch_{c} away from them, see [9, 10] for a more detailed description.

Clearly, we first solve the linear systems (6.2a) in order to obtain 𝐮¯hn{\overline{{\bf u}}}_{h}^{n}, then (6.2b) for 𝐮^hn{\widehat{{\bf u}}}_{h}^{n}, then (6.2c) for 𝐪^hn\widehat{{\bf q}}_{h}^{n}, then (6.2d) for 𝒑¯hn\overline{\bm{p}}_{h}^{n}, and finally the variational inequality (6.2e) for φhn\varphi_{h}^{n}. In two space dimensions we employ the package LDL, see [23], together with the sparse matrix ordering AMD, see [6], in order to solve the linear systems (6.2a)–(6.2d). In three space dimensions, on the other hand, we use a preconditioned conjugate gradient algorithm, with a WW-cycle multigrid step as preconditioner. In order to solve the variational inequality (6.2e) we employ a nonlinear multigrid method similar to [42].

For the computational domain we will choose Ω¯=[0,L1]×[12,12]\overline{\Omega}=[0,L_{1}]\times[-\frac{1}{2},\frac{1}{2}] in two dimensions, and Ω¯=[0,L1]×[12L2,12L2]×[12,12]\overline{\Omega}=[0,L_{1}]\times[-\frac{1}{2}L_{2},\frac{1}{2}L_{2}]\times[-\frac{1}{2},\frac{1}{2}] in three dimensions with positive lengths LiL_{i}, i{1,2}i\in\{1,2\}, given below. For the physical parameters we loosely follow the settings in [50]. In particular, for the forcings we choose 𝐅¯=𝐅^=𝟎\overline{{\bf F}}=\widehat{{\bf F}}={\bf 0} throughout, as well as 𝒈^=𝟎\widehat{\bm{g}}={\bf 0} and

𝒈¯(𝒙)={g𝒆1x1=L1,𝟎x1<L1,withg=0.1.\overline{\bm{g}}(\bm{x})=\begin{cases}g{\bm{e}}_{1}&x_{1}=L_{1},\\ {\bf 0}&x_{1}<L_{1},\end{cases}\qquad\text{with}\quad g=0.1. (6.3)

For the interpolated elasticity tensors we choose ¯(s)=12[(1+s)¯++(1s)¯]\overline{{\mathbb{C}}}(s)=\frac{1}{2}[(1+s)\overline{{\mathbb{C}}}_{+}+(1-s)\overline{{\mathbb{C}}}_{-}], where the two tensors ¯±\overline{{\mathbb{C}}}_{\pm} are defined through the Young’s moduli E¯±\overline{E}_{\pm} and Poisson ratios ν¯±\overline{\nu}_{\pm} via

E¯+=3,E¯=0.7,ν¯+=ν¯=0.45.\overline{E}_{+}=3,\ \overline{E}_{-}=0.7,\quad\overline{\nu}_{+}=\overline{\nu}_{-}=0.45. (6.4)

Similarly, ^(s)=12[(1+s)^++(1s)^]\widehat{{\mathbb{C}}}(s)=\frac{1}{2}[(1+s)\widehat{{\mathbb{C}}}_{+}+(1-s)\widehat{{\mathbb{C}}}_{-}], with the Young’s moduli and Poisson ratios

E^+=13,E^=0.6,ν^+=ν^=0.45.\widehat{E}_{+}=13,\ \widehat{E}_{-}=0.6,\quad\widehat{\nu}_{+}=\widehat{\nu}_{-}=0.45. (6.5)

For the fixity function we choose

χ(s)=25(1+s).\chi(s)=\tfrac{2}{5}(1+s). (6.6)

For the visualisation of the progress of the discrete gradient flow computations, we define the discrete cost functional

Jh(φhh,𝐮^hn)\displaystyle J^{h}(\varphi_{h}^{h},{\widehat{{\bf u}}}_{h}^{n}) =γ(ε2|φhn|2+1εψ(φhn),1)+12ΓtarW(𝐮^hn𝒖tar)(𝐮^hn𝒖tar)dd1\displaystyle=\gamma\left(\frac{\varepsilon}{2}|\nabla\varphi_{h}^{n}|^{2}+\frac{1}{\varepsilon}\psi(\varphi_{h}^{n}),1\right)+{\frac{1}{2}}\int_{\Gamma^{\rm tar}}W({\widehat{{\bf u}}}_{h}^{n}-\bm{u}^{\rm tar})\cdot({\widehat{{\bf u}}}_{h}^{n}-\bm{u}^{\rm tar})\,\mathrm{d}\mathcal{H}^{d-1}
=:γh(φhn)+h,tar(𝐮^hn).\displaystyle=:\gamma\mathcal{E}^{h}(\varphi_{h}^{n})+\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}). (6.7)

As the initial data φh0\varphi_{h}^{0} we choose a random mixture with mean zero. Choosing other initial data, including random mixtures with positive or negative mean had no visible influence on the obtained optimal shapes.

6.2 Numerical simulations in two dimensions

For the target shapes we consider the parabolic profile

𝒖tar(x1,x2)=utar(x1)𝒆2,utar(x1)=ctar(x1)2,\bm{u}^{\rm tar}(x_{1},x_{2})=u^{\rm tar}(x_{1}){\bm{e}}_{2},\quad u^{\rm tar}(x_{1})=c^{\rm tar}(x_{1})^{2}, (6.8a)
with ctar>0c^{\rm tar}>0, and the cosine profile
𝒖tar(x1,x2)=utar(x1)𝒆2,utar(x1)=ctar(1cosktarπx1L1)\bm{u}^{\rm tar}(x_{1},x_{2})=u^{\rm tar}(x_{1}){\bm{e}}_{2},\quad u^{\rm tar}(x_{1})=c^{\rm tar}\Big{(}1-\cos\frac{k^{\rm tar}\pi x_{1}}{L_{1}}\Big{)} (6.8b)

with ctar>0c^{\rm tar}>0 and ktar>0k^{\rm tar}>0. In Figure 2 we plot two examples for the profiles in (6.8) for the domain length L1=12L_{1}=12.

In each of the Figures 3, 4, 5 and 6, we provide visualisations of the numerical solution φhn\varphi_{h}^{n} at various pseudo-times (black denotes the passive material {φhn=1}\{\varphi_{h}^{n}=-1\} and grey denotes the active material {φhn=1}\{\varphi_{h}^{n}=1\}), and the corresponding displacements 𝐮¯hn{\overline{{\bf u}}}_{h}^{n} (in red) and 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (in green). As mentioned before, each time the gradient flow is started from a random mixture φh0\varphi_{h}^{0} with zero mean. We also provide pseudo-time plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion of the elastic h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) and interfacial γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) energies, as well as log-plots of the elastic energies. The parameter details are summarised in the Table 1.

Figure Profile Domain WW Γtar\Gamma^{\rm tar} ctarc^{\rm tar} ktark^{\rm tar}
3 Parabolic (6.8a) [0,6]×[12,12][0,6]\times[-\frac{1}{2},\frac{1}{2}] Id\mathrm{Id} Ω\partial\Omega 0.0750.075 -
4 Cosine (6.8b) [0,6]×[12,12][0,6]\times[-\frac{1}{2},\frac{1}{2}] 𝒆2𝒆2\bm{e}_{2}\otimes\bm{e}_{2} Ω\partial\Omega 0.250.25 22
5 Parabolic (6.8a) [0,12]×[12,12][0,12]\times[-\frac{1}{2},\frac{1}{2}] Id\mathrm{Id} Ω\partial\Omega 0.020.02 -
6 Cosine (6.8b) [0,12]×[12,12][0,12]\times[-\frac{1}{2},\frac{1}{2}] Id\mathrm{Id} topΩ\partial_{\rm top}\Omega 11 1.51.5
Table 1: Parameter details for numerical simulations in 2D.

For all the presented simulations we choose the parameters ε=18π\varepsilon=\frac{1}{8\pi} and γ=0.01\gamma=0.01. In each case the cost functional decreases monotonically, but the proportions of the two energies (elastic vs interfacial) differ from case to case.

The first experiment in Figure 3 is for the parabolic profile on the domain [0,6]×[12,12][0,6]\times[-\frac{1}{2},\frac{1}{2}]. We observe that in the optimal distribution of material, the active phase occupies most of the lower domain. This ensures that in the programmed stage, the printed active composite is able to attain the desired target shape.

Refer to caption
Refer to caption
Figure 2: Plots of the function utaru^{\rm tar} in (6.8a), with ctar=0.02c^{\rm tar}=0.02, (left) and (6.8b), with ctar=1c^{\rm tar}=1 and ktar=1.5k^{\rm tar}=1.5, (right), when L1=12L_{1}=12.

Not surprisingly, a very different distribution of material is obtained when changing the target shape functional to enforce a cosine profile at the programmed stage. As can be seen from Figure 3, the optimal distribution of the active material is now given by an elongated region that connects the lower left of the domain with the upper right.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Computation on Ω¯=[0,6]×[12,12]\overline{\Omega}=[0,6]\times[-\tfrac{1}{2},\tfrac{1}{2}] for the target shape (6.8a) with ctar=0.075c^{\rm tar}=0.075 and W=IdW=\mathrm{Id} and Γtar=Ω\Gamma^{\rm tar}=\partial\Omega, with ε=18π\varepsilon=\frac{1}{8\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} and the displacements 𝐮¯hn{\overline{{\bf u}}}_{h}^{n} (red) and 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (green) at pseudo-times t=0.1, 0.5, 1t=0.1,\,0.5,\,1. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

It turns out that on longer (or thinner) domains, far less active material is needed to achieve significant deformations at the programmed state. For example, in Figure 4 a miniscule amount of active material, spread in several connected components arranged at the bottom of the domain, is sufficient to result in the desired parabolic target shape.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption

Figure 4: Computation on Ω¯=[0,6]×[12,12]\overline{\Omega}=[0,6]\times[-\tfrac{1}{2},\tfrac{1}{2}] for the target shape (6.8b) with ctar=0.25c^{\rm tar}=0.25, ktar=2k^{\rm tar}=2 and W=𝒆2𝒆2W=\bm{e}_{2}\otimes\bm{e}_{2} and Γtar=Ω\Gamma^{\rm tar}=\partial\Omega, with ε=18π\varepsilon=\frac{1}{8\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} and the displacements 𝐮¯hn{\overline{{\bf u}}}_{h}^{n} (red) and 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (green) at pseudo-times t=1, 2, 5t=1,\,2,\,5. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

Similarly, we observe in Figure 5 that three strategically placed small amounts of active material guarantee a cosine profile at the programmed stage for the printed active composite.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Computation on Ω¯=[0,12]×[12,12]\overline{\Omega}=[0,12]\times[-\tfrac{1}{2},\tfrac{1}{2}] for the target shape (6.8a) with ctar=0.02c^{\rm tar}=0.02 and W=IdW=\mathrm{Id} and Γtar=Ω\Gamma^{\rm tar}=\partial\Omega, with ε=18π\varepsilon=\frac{1}{8\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} and the displacements 𝐮¯hn{\overline{{\bf u}}}_{h}^{n} (red) and 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (green) at pseudo-times t=0.5, 1, 2t=0.5,\,1,\,2. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Computation on Ω¯=[0,12]×[12,12]\overline{\Omega}=[0,12]\times[-\tfrac{1}{2},\tfrac{1}{2}] for the target shape (6.8b) with ctar=1c^{\rm tar}=1, ktar=1.5k^{\rm tar}=1.5 and W=IdW=\mathrm{Id} and Γtar=topΩ\Gamma^{\rm tar}=\partial_{\rm top}\Omega, with ε=18π\varepsilon=\frac{1}{8\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} and the displacements 𝐮¯hn{\overline{{\bf u}}}_{h}^{n} (red) and 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (green) at pseudo-times t=0.1, 0.5, 1t=0.1,\,0.5,\,1. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

6.3 Numerical simulations in three dimensions

For the target shapes we consider a function of the form

𝒖tar(x1,x2,x3)=utar(x1,x2)𝒆3\bm{u}^{\rm tar}(x_{1},x_{2},x_{3})=u^{\rm tar}(x_{1},x_{2}){\bm{e}}_{3}

and choose utaru^{\rm tar} from one of the following:

utar(x1,x2)=ctar(x1)2.u^{\rm tar}(x_{1},x_{2})=c^{\rm tar}(x_{1})^{2}. (6.9)
utar(x1,x2)=ctar(1cosktarπx1L1).u^{\rm tar}(x_{1},x_{2})=c^{\rm tar}\Big{(}1-\cos\frac{k^{\rm tar}\pi x_{1}}{L_{1}}\Big{)}. (6.10)
utar(x1,x2)=ctarx1x2.u^{\rm tar}(x_{1},x_{2})=c^{\rm tar}x_{1}x_{2}. (6.11)

Notice that (6.9) and (6.10) are simply the three dimensional analogues of the parabolic profile (6.8a) and the cosine profile (6.8b), respectively. On the other hand, (6.11) yields a linear profile in x1x_{1} with a twisting in the x2x_{2}-direction.

In each of the Figures 7, 8, and 9, 10 we provide visualisations of the numerical solution φhn\varphi_{h}^{n} at various pseudo-times (black denotes the passive material {φhn=1}\{\varphi_{h}^{n}=-1\} and grey denotes the active material {φhn=1}\{\varphi_{h}^{n}=1\}), and the corresponding displacements 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (darker colours indicating lower values of 𝐮^hn𝒆3{\widehat{{\bf u}}}_{h}^{n}\cdot\bm{e}_{3} and lighter colours indicating higher values of 𝐮^hn𝒆3{\widehat{{\bf u}}}_{h}^{n}\cdot\bm{e}_{3}). We also provide pseudo-time plots of the energy functionals, similarly to the 2D simulations in the previous subsection. The parameter details are summarised in the Table 2.

Figure Profile Domain WW Γtar\Gamma^{\rm tar} ctarc^{\rm tar} ktark^{\rm tar}
7 (6.9) [0,6]×[32,32]×[0,1][0,6]\times[-\frac{3}{2},\frac{3}{2}]\times[0,1] Id\mathrm{Id} topΩ\partial_{\rm top}\Omega 0.0750.075 -
8 (6.10) [0,12]×[32,32]×[0,1][0,12]\times[-\frac{3}{2},\frac{3}{2}]\times[0,1] 𝒆3𝒆3\bm{e}_{3}\otimes\bm{e}_{3} topΩ\partial_{\rm top}\Omega 1 2
9 (6.10) [0,12]×[32,32]×[0,1][0,12]\times[-\frac{3}{2},\frac{3}{2}]\times[0,1] Id\mathrm{Id} topΩ\partial_{\rm top}\Omega 11 22
10 (6.11) [0,6]×[3,3]×[0,1][0,6]\times[-3,3]\times[0,1] 𝒆3𝒆3\bm{e}_{3}\otimes\bm{e}_{3} rightΩ\partial_{\rm right}\Omega 0.10.1 -
Table 2: Parameter details for numerical simulations in 3D.

In the first three figures we take ε=14π\varepsilon=\frac{1}{4\pi} and for γ\gamma choose either 0.10.1 or 0.010.01. In each simulation the cost functional decreases monotonically, but the proportions of the two energies (elastic vs interfacial) differ from case to case.

The first simulation is a direct 3D analogue for the computation previously shown in Figure 3, the only difference being that here we restrict the set Γtar\Gamma^{\rm tar} to the upper part of the boundary Ω\partial\Omega. As expected, the observed results are very close to the ones seen previously in the 2D setting. In particular, the active phase occupies the lower half of the domain, with a dip towards the right end of the domain.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Computation on Ω¯=[0,6]×[32,32]×[0,1]\overline{\Omega}=[0,6]\times[-\tfrac{3}{2},\tfrac{3}{2}]\times[0,1] for the target shape (6.9) with ctar=0.075c^{\rm tar}=0.075 and W=IdW=\mathrm{Id} and Γtar=topΩ\Gamma^{\rm tar}=\partial_{\rm top}\Omega, with ε=14π\varepsilon=\frac{1}{4\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} (side view and bottom view) and the displacement 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (with colour coding for 𝐮^hn𝒆3{\widehat{{\bf u}}}_{h}^{n}\cdot\bm{e}_{3}) at pseudo-times t=1, 2, 5t=1,\,2,\,5. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

On more elongated domains we once again observe that relatively little active material can result in large deformations at the programmed stage. For example, the strategic placement of the active component seen in Figure 8 yields a large cosine profile deformation.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: Computation on Ω¯=[0,12]×[32,32]×[0,1]\overline{\Omega}=[0,12]\times[-\tfrac{3}{2},\tfrac{3}{2}]\times[0,1] for the target shape (6.10) with ctar=1c^{\rm tar}=1, ktar=2k^{\rm tar}=2 and W=𝒆3𝒆3W=\bm{e}_{3}\otimes\bm{e}_{3} and Γtar=topΩ\Gamma^{\rm tar}=\partial_{\rm top}\Omega, with ε=14π\varepsilon=\frac{1}{4\pi} and γ=0.1\gamma=0.1. We display φhn\varphi_{h}^{n} (side view and bottom view) and the displacement 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (with colour coding for 𝐮^hn𝒆3{\widehat{{\bf u}}}_{h}^{n}\cdot\bm{e}_{3}) at pseudo-times t=1, 2, 10t=1,\,2,\,10. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

It is interesting to note that by simply changing the weighting matrix WW in the target energy functional, we obtain a completely different optimal distribution of active material. This can be seen in Figure 9, where the only change to the previous simulation is W=IdW=\mathrm{Id}, rather than W=𝒆3𝒆3W=\bm{e}_{3}\otimes\bm{e}_{3}. Now there are just two connected components for the active phase, one at the lower left part of the domain, and one at the middle of the top of the domain. Yet the obtained deformations at the programmed stage are very similar.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Computation on Ω¯=[0,12]×[32,32]×[0,1]\overline{\Omega}=[0,12]\times[-\tfrac{3}{2},\tfrac{3}{2}]\times[0,1] for the target shape (6.10) with ctar=1c^{\rm tar}=1, ktar=2k^{\rm tar}=2 and W=IdW=\mathrm{Id} and Γtar=topΩ\Gamma^{\rm tar}=\partial_{\rm top}\Omega, with ε=14π\varepsilon=\frac{1}{4\pi} and γ=0.01\gamma=0.01. We display φhn\varphi_{h}^{n} (side view and bottom view) and the displacement 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (with colour coding for 𝐮^hn𝒆3{\widehat{{\bf u}}}_{h}^{n}\cdot\bm{e}_{3}) at pseudo-times t=0.1, 0.5, 1t=0.1,\,0.5,\,1. Below we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}).

Our final numerical simulation is for a twisted target shape. In particular, we use the target function (6.11) with ctar=0.1c^{\rm tar}=0.1 on the domain Ω¯=[0,6]×[3,3]×[0,1]\overline{\Omega}=[0,6]\times[-3,3]\times[0,1], with W=𝒆3𝒆3W=\bm{e}_{3}\otimes\bm{e}_{3} and Γtar=rightΩ\Gamma^{\rm tar}=\partial_{\rm right}\Omega. In Figure 10 we provide visualisations of the numerical solution φhn\varphi_{h}^{n} at various pseudo-times, and the corresponding displacements 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (here darker colours indicate lower values of |𝐮^hn||{\widehat{{\bf u}}}_{h}^{n}| and lighter colours indicate higher values of |𝐮^hn||{\widehat{{\bf u}}}_{h}^{n}|). For this computation we take ε=12π\varepsilon=\frac{1}{2\pi} and γ=0.02\gamma=0.02. At the optimal configuration we observe an elaborate distribution of the active material, which yields a twisted shape of the component at the programmed stage.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Computation on Ω¯=[0,6]×[3,3]×[0,1]\overline{\Omega}=[0,6]\times[-3,3]\times[0,1] for the target shape (6.11)(b) with ctar=0.1c^{\rm tar}=0.1 and W=𝒆3𝒆3W=\bm{e}_{3}\otimes\bm{e}_{3} and Γtar=rightΩ\Gamma^{\rm tar}=\partial_{\rm right}\Omega, with ε=12π\varepsilon=\frac{1}{2\pi} and γ=0.02\gamma=0.02. We display φhn\varphi_{h}^{n} and the displacement 𝐮^hn{\widehat{{\bf u}}}_{h}^{n} (with colour coding for |𝐮^hn||{\widehat{{\bf u}}}_{h}^{n}|) at pseudo-times t=0.1, 1, 2t=0.1,\ 1,\,2. In the middle we show plots of the cost functional Jh(φhn,𝐮^hn)J^{h}(\varphi_{h}^{n},{\widehat{{\bf u}}}_{h}^{n}), the proportion in it of the elastic energy h,tar(𝐮^hn)\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}) (solid blue) and the interfacial energy γh(φhn)\gamma\mathcal{E}^{h}(\varphi_{h}^{n}) (dashed red), as well as of log10h,tar(𝐮^hn)\log_{10}\mathcal{E}^{\rm h,tar}({\widehat{{\bf u}}}_{h}^{n}). In the bottom we show a backward view of the first row of plots.

Acknowledgments

The authors HG and AS gratefully acknowledge the support by the Graduiertenkolleg 2339 IntComSin of the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 321821685. The work of KFL is supported by the Research Grants Council of the Hong Kong Special Administrative Region, China [Project No.: HKBU 14302218 and HKBU 12300321].

References

  • [1] S.A. Abdullah, A. Jumahat, N.R. Abdullah and L. Frormann. Setermination of Shape Fixity and Shape Recovery Rate of Carbon Nanotube-flled Shape Memory Polymer Nanocomposites. Procedia Eng. 41 (2012) 1641–1646
  • [2] M.H. Ali, A. Abilgaziyev and D. Adair. 4D printing: a critical review of current developments, and future prospects. Int. J. Adv. Manuf. Technol. 104 (2019) 701–717
  • [3] G. Allaire and B. Bogosel. Optimizing supports for additive manufacturing. Struct. Multidiscip. Optim. 58 (2018) 2493–2515
  • [4] S. Almi and U. Stefanelli. Topology optimization for incremental elastoplasticity: A phase-field approach. SIAM J. Control Optim. 59 (2021) 339–364
  • [5] H.W. Alt: Linear Functional Analysis, an Application Oriented Introduction. Springer, London, (2016)
  • [6] P.R. Amestoy, T.A. Davis and I.S. Duff. Algorithm 837: AMD, an approximate minimum degree ordering algorithm. ACM Trans. Math. Software 30 (2004) 381–388
  • [7] S.E. Bakarich, E. Gorkin and G.M. Spinks. 4D printing with mechanically robust, thermally actuating hydrogels. Macromol. Rapid Commun. 36 (2015) 1211–1217
  • [8] S. Baldo. Minimal interface criterion for phase transitions in mixtures of Cahn–Hilliard fluids. Ann. Inst. H. Poincaré Anal. Non Linéaire 7 (1990) 67–90
  • [9] L. Baňas and R. Nürnberg. Finite element approximation of a three dimensional phase field model for void electromigration. J. Sci. Comp. 37 (2008) 202–232
  • [10] J.W. Barrett, R. Nürnberg and V. Styles. Finite element approximation of a phase field model for void electromigration. SIAM J. Numer. Anal. 42 (2004) 738–772
  • [11] G. Bellettini, A. Braides and G. Riey. Variational approximation of anisotropic functionals on partitions. Annali di Matematica 184 (2005) 75–93
  • [12] A. Bhattacharyya and K.A. James. Topology optimization of shape memory polymers structures with programmable morphology. Struct. Multidiscip. Optim. 63 (2021) 1863–1887
  • [13] L. Blank, H. Garcke, M. Hassan Farshbaf-Shaker and V. Styles. Relating phase field and sharp interface approaches to structural topology optimization. ESAIM: Control Optim. Calc. Var. 20 (2014) 1025–1058
  • [14] L. Blank, H. Garcke, C. Hecht and C. Rupprecht. Sharp interface limit for a phase field model in structural optimization. SIAM J. Control Optim. 54 (2016) 1558–1584
  • [15] L. Blank, H. Garcke, L. Sarbu, T. Srisupattarawanit, V. Styles and A. Voigt. Phase-field approaches to structural topology optimization, in Constrained optimization and optimal control for partial differential equations, vol. 160 of Internat. Ser. Numer. Math., Birkhäuser/Springer Basel AG, Basel, 2012, 245–256
  • [16] J.F. Blowey and C.M. Elliott. Curvature dependent phase boundary motion and parabolic double obstacle problems, in Degenerate diffusions (Minneapolis, MN, 1991), vol. 47 of IMA Vol. Math. Appl., Springer, New York, 1993, 19–60
  • [17] B. Bourdin and A. Chambolle. Design-dependent loads in topology optimization. ESAIM: Control Optim. Cal. Var. 9 (2003) 19–48
  • [18] L. Bronsard, H. Garcke and B. Stoth. A multi-phase Mullins-Sekerka system: matched asymptotic expansions and an implicit time discretization for the geometric evolution problem. SIAM J. Appl. Math. 60 (1999) 295–315
  • [19] M. Burger and R. Stainko. Phase-field relaxation of topology optimization with local stress constraints. SIAM J. Control Optim. 45 (2006) 1447–1466
  • [20] S. Cacace, E. Cristiani and L. Rocchi. A Level Set Based Method for Fixing Overhangs in 3D Printing. Appl. Math. Model. 44 (2017) 446–455
  • [21] M. Carraturo, E. Rocca, E. Bonetti, D. Hömberg, A. Reali and F. Auricchio. Graded-material design based phase-field and topology optimization. Comput. Mech. 64 (2019) 1589–1600
  • [22] P.G. Ciarlet: The Finite Element Method for Elliptic Problems. North-Holland Publishing Co., Amsterdam, (1978)
  • [23] T.A. Davis, Algorithm 849: a concise sparse Cholesky factorization package. ACM Trans. Math. Software 31 (2005) 587–591
  • [24] Z. Ding, C. Yuan, X. Peng, T. Wang, JH.J. Qi and M.L. Dunn. Direct 4D printing via active composite materials. Sci. Adv. 3 (2017) e1602890
  • [25] P.C. Fife and O. Penrose. Interfacial dynamics for thermodynamically consistent phase-field models with nonconserved order parameter. J. Differential Equations 16 (1995) 1–49
  • [26] J. Jiang, X. Xu and J. Stringer. Support Structures for Additive Manufacturing: A Review. J. Mauf. Mater. Process. 2 (2008) 64 (23 pages)
  • [27] H. Garcke. The Γ\Gamma-limit of the Ginzburg–Landau energy in an elastic medium. AMSA 18 (2008) 345–379
  • [28] H. Garcke and C. Hecht. Apply a phase field approach for shape optimization of a stationary Navier-Stokes flow. ESAIM: Control Optim. Cal. Var. 22 (2016) 309–337
  • [29] H. Garcke and C. Hecht. Shape and topology optimization in Stokes flow with a phase field approach. Appl. Math. Optim. 73 (2016) 23–70
  • [30] H. Garcke, C. Hecht, M. Hinze, C. Kahle and K.F. Lam. Shape optimization for surface functionals in Navier–Stokes flow using a phase field approach. Interfaces Free Bound. 18 (2016) 219–261
  • [31] H. Garcke, P. Hüttl and P. Knopf. Shape and topology optimization involving the eigenvalues of an elastic structure: A multi-phase-field approach. Adv. Nonlinear Anal. 11 (2022) 159–197
  • [32] H. Garcke, K.F. Lam, R. Nürnberg and A. Signori. Overhang penalization in additive manufacturing via phase field structural topology optimization with anisotropic energies. Preprint ArXiv:2111.14070
  • [33] H. Garcke, B. Nestler and B. Stoth. On anisotropic order parameter models for multi0phase systems and their sharp interface limits. Phys. D 115 (1998) 87–108
  • [34] H. Garcke and B. Stinner. Second order phase field asymptotics for multi-component systems. Interfaces Free Bound. 8 (2006) 131–157
  • [35] Q. Ge, C.K. Dunn, H.J. Qi and M.L. Dunn. Active origami by 4D printing. Smart Mater. Struct. 23 (2014) 094007
  • [36] O. Kuksenok and A.C. Balazs. Stimuli-responsive behavior of composites integating thermo-responsive gels with photo-responsive fibers. Mater. Horizons 3 (2016) 53–62
  • [37] Q. Ge, H.J. Qi and M.L. Dunn. Active materials by four-dimension printing. Appl. Phys. Lett. 103 (2013) 131901
  • [38] Q. Ge, A.H. Sakhaei, H. Lee, C.K. Dunn, N.X. Fang and M.L. Dunn. Multimaterial 4D Printing with Tailorable Shape Memory Polymers. Sci. Rep. 6 (2016) 31110
  • [39] M.J. Geiss, N. Boddeti, O. Weeger, K. Maute and M.L. Dunn. Combined level-set-XFEM-density topology optimization of four-dimensional printed structures undergoing large deformation. J. Mech. Des. 141 (2019) 051405
  • [40] M. Howard, J. Pajot, K. Maute and M.L. Dunn. A Computational Design Methodology for Assembly and Actuation of Thin-Film Structures via Patterning of Eigenstrains. J. Microelectromech. Syst. 18 (2009) 1137–1148
  • [41] C. Hecht. Shape and topology optimization in fluids using a phase field approach an application in structural optimization. PhD Thesis University of Regensburg, Regensburg, Germany 2014 https://epub.uni-regensburg.de/29869/1/Dissertation_ClaudiaHecht.pdf
  • [42] R. Kornhuber. Monotone multigrid methods for elliptic variational inequalities II. Numer. Math. 72 (1996) 481–499
  • [43] M. Leray, L. Merli, F. Torti, M. Mazur and M. Bandt. Optimal topology for additive manufacturing: A method for enabling additive manufacture of support-free optimal structures. Mater. Des. 63 (2014) 678–690
  • [44] J. Liu, A.T. Gaynor, S. Chen, Z. Kang, K. Suresh, A. Takezawa, L. Li, J. Kato, J. Tang, C.C.L. Wang, L. Cheng, X. Liang and A.C. To. Current and future trends in topology optimization for additive manufacturing. Struct. Multi. Optim. 57 (2018) 2457–2483
  • [45] J. Liu and H. Yu. Self-Supporting Topology Optimization With Horizontal Overhangs for Additive Manufacturing. J. Manuf. Sci. Eng. 142 (2020) 091003 (14 pages)
  • [46] M. Langelaar. Combined optimization of part topology, support structure layout and build orientation for additive manufacturing. Struct. Multidiscip. Optim. 57 (2018) 1985–2004
  • [47] H. Li, X. Gao and Y. Luo. Multi-shape memory polymers achieved by the spatio-assembly of 3D printable thermoplastic building blocks. Soft Matter 12 (2016) 3226–3233
  • [48] J.L. Lions and E. Magenes. Non-homogeneous boundary value problems and applications. Springer, 1972.
  • [49] T.S. Lumpe and K. Shea. Computational design of 3D-printed active lattice structures for reversible shape morphing. J. Mater. Res. 36 (2021) 3642–3655
  • [50] K. Maute, A. Tkachuk, J. Wu, H.J. Qi, Z. Ding and M.L. Dunn. Level Set Topology Optimization of Printed Active Composites. J. Mech. Des. 137 (2015) 111402 (13 pages)
  • [51] A.M. Mirzendehdel and K. Suresh. Support Structure Constrained Topology Optimization for Additive Manufacturing. Comput. Aided Des. 1 (2016) 1–13
  • [52] F. Momeni, S.M. Hassani, N.X. Liu and J. Ni. A review of 4D printing. Mater. Des. 122 (2017) 42–79
  • [53] N.C. Owen, J. Rubinstein, P. Sternberg. Minimizers and gradient flows for singularly perturbed bi-stable potentials with a Dirichlet condition. Proc. R. Soc. Lond. A 429 (1990) 505–532
  • [54] J.M. Pajot, K. Maute, Y. Zhang and M.L. Dunn. Design of patterned multilayer films with eigenstrains by topology optimization. Int. J. Solids Struct. 43 (2006) 1832–1853
  • [55] P. Penzler, M. Rumpf and B. Wirth. A phase-field model for compliance shape optimization in nonlinear elasticity. ESAIM Control Optim. Cal. Var. 18 (2012) 229–258
  • [56] A. Schmidt and K.G. Siebert. Design of Adaptive Finite Element Software: The Finite Element Toolbox ALBERTA. vol. 42 of Lecture Notes in Computational Science and Engineering, Springer-Verlag, Berlin, 2005
  • [57] J. Sokolowski and J.-P. Zolesio. Introduction to Shape Optimization. Springer series in computational mathematics, Vol. 16, Springer-Verlag, Berlin Heidelberg, 1991
  • [58] L. Sun and W.M. Huang. Mechanisms of the multi-shape memory effect and temperature memory effect in shape memory polymers. Soft Matter 6 (2010) 4403–4406
  • [59] Y. Sun, W. Ouyang, Z. Liu, N. Ni, Y. Savoye, P. Song and L. Liu. Computational Design of Self-actuated Deformable Solids via Shape Memory Material. IEEE Trans. Vis. Comput. Graph. —bf 28 (2022) 2577–2588
  • [60] A. Takezawa, S. Nishiwaki and M. Kitamura. Shape and topology optimization based on the phase field method and sensitivity analysis. J. Comput. Phys. 229 (2010) 2697–2718
  • [61] S. Tibbits. 4D printing: Multi-material shape change. Archit. Des. 84 (2014) 116–121
  • [62] M.Y. Wang and S. Zhou. Phase field: a variational method for structural topology optimization. CMES Comput. Model. Eng. Sci. 6 (2004) 547–566
  • [63] X. Wan, Y. He, Y. Liu and J. Leng. 4D printing of multiple shape memory polymer and nanocomposites with biocompatible programmable and selectively actuated properties. Addit. Manuf. 53 (2022) 102689
  • [64] G. Wang, H. Yang, Z. Yan, N.G. Ulu, Y. Tao, J. Gu, L.B. Kara and L. Yao. 4DMesh: 4d printing morphing non-developable mesh surfaces. in the 31st Annual ACM Symposium on User Interface Software and Technology (2018) 623–635
  • [65] Y. Wei, P. Huang, Z. Li, P. Wang and X. Feng. Design of active materials distributions for four-dimensional printing based on multi-material topology optimization. Smart Mater. Struct. 30 (2021) 095002
  • [66] J. Wu, C. Yuan, Z. Ding, M. Isakov, Y. Mao, T. Wang, M.L. Dunn and H.J. Qi. Multi-shape active composites by 3D printing of digital shape memory polymers. Sci. Rep. 6 (2016) 24224
  • [67] C. Yuan, T. Lu and T.J. Wang. Mechanics-based design strategies for 4D printing: A review. Forces Mech. 7 (2022) 100081.
  • [68] Q. Zhang, K. Zhang and G. Hu. Smart three-dimensional lightweight structure triggered from a thin composite sheet via 3D printing technique. Sci. Rep. 6 (2016) 22431