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

Overhang penalization in additive manufacturing via phase field structural topology optimization with anisotropic energies

Harald Garcke 111Fakultät für Mathematik, Universität Regensburg, 93040 Regensburg, Germany ([email protected]).    Kei Fong Lam 222Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong ([email protected]).    Robert Nürnberg 333Department of Mathematics, University of Trento, Trento, Italy ([email protected]).    Andrea Signori 444Dipartimento di Matematica “F. Casorati”, Università di Pavia, 27100 Pavia, Italy ([email protected]).
Abstract

A phase field approach for structural topology optimization with application to additive manufacturing is analyzed. The main novelty is the penalization of overhangs (regions of the design that require underlying support structures during construction) with anisotropic energy functionals. Convex and non-convex examples are provided, with the latter showcasing oscillatory behavior along the object boundary termed the dripping effect in the literature. We provide a rigorous mathematical analysis for the structural topology optimization problem with convex and non-continuously-differentiable anisotropies, deriving the first order necessary optimality condition using subdifferential calculus. Via formally matched asymptotic expansions we connect our approach with previous works in the literature based on a sharp interface shape optimization description. Finally, we present several numerical results to demonstrate the advantages of our proposed approach in penalizing overhang developments.

Keywords: Topology optimization, phase field, anisotropy, linear elasticity, optimal control, additive manufacturing, overhang penalization

AMS (MOS) Subject Classification: 49J20, 49K40, 49J50

1 Introduction

Additive manufacturing (AM) is an innovative building technique that produces objects in a layer-by-layer fashion through fusing or binding raw materials in powder and resin forms. Since its introduction in the 1970s, it has shown great versatility in allowing for the creation of highly complex geometries, immediate modifications and redesign, thus making it an ideal process for rapid prototyping and testing. But despite such advantages over traditional manufacturing technologies, AM still has not seen widespread integration in serial production, and there are still many limitations that have yet to be overcome. For an overview of the technologies involved and the challenges encountered by AM, we refer the reader to the review article [1] as well as the report [74].

In this work, we focus on a recurring issue encountered when practitioners employ AM to construct objects. Some categories of AM technologies, such as Fused Deposition Modeling and Laser Metal Deposition, construct objects in an upwards direction (hereafter referred to as the build direction) by repeatedly depositing and then fusing a new layer of raw materials on the surface of the object. Overhangs are regions of the constructed object that when placed in a certain orientation extend outwards without any underlying support. For example, the top horizontal bar of a T-shaped structure placed vertically will be classified as an overhang. The main issue with overhangs is that they can deform under their own weight or by thermal residual stresses from the construction process, and if not supported from below, there is a risk that the object itself can display unintended deformations and the entire printing process can fail.

One natural solution is to identify certain orientations of the object whose overhang regions are minimized prior to printing, and choose the build direction to be one of these orientations. Going back to the T-shape structure example, we can simply rotate the shape by 180180^{\circ} so that there are no overhang regions (see, e.g., [54, Fig. 5]). However, in general, there is no guarantee that orientations with no overhang regions exist. In this direction we mention the works [70, 91] that incorporate various geometric information such as contact surface area and support volume expressed as functions of the build direction within an optimization procedure. On the other hand, simultaneous optimization of build direction and topology has been considered in [88].

Another remedy is to employ support structures that are built concurrently with the object whose purpose is to reduce the potential deformation of overhang regions through mechanical loads or thermal residual stresses. These supplementary structures act like scaffolding to overhangs, and after a successful print are then removed in a post-processing step. Along with increased material costs and printing time, there is a risk of damaging delicate features of the objects during the removal step. Nevertheless, for certain AM technologies such as the aforementioned Fused Deposition Modeling, some form of support structures will always be necessary in order to mitigate the potential undesirable deformations of the finished object. Thus, recent mathematical research has concentrated on the optimizing support structures in an effort to reduce material waste and minimize contact area with the object. Many such works explore optimal support in the framework of shape and topology optimization [3, 48, 56, 60, 67], as well as proposing sparse cellular designs that have low solid volume and contact area in the form of lattices [53], honeycomb [64], and trees [87]. Further details can be found in the review article [54].

A related approach would be to allow some modifications to the object, as long as the altered design retains the intended functionality of the original, leading to creation of self-supporting objects that do not require support structures at all [30, 61, 62, 63]. Our present work falls roughly into this category, where we employ a well-known phase field methodology in structural topology optimization with the aim of identifying optimal designs fulfilling the so-called overhang angle constraint. In mathematical terms, let us consider an object Ω1d\Omega_{1}\subset\mathbb{R}^{d}, d=2,3,d=2,3, being built within a hold-all rectangular domain Ω=[0,1]d\Omega=[0,1]^{d} in a layer-by-layer fashion with build direction 𝒆d=(0,,1)\bm{e}_{d}=(0,\dots,1)^{\top}. The base plate is the region :=[0,1]d1×{0}\mathcal{B}:=[0,1]^{d-1}\times\{0\} where Ω1\Omega_{1} is assembled on. For any 𝒑Ω1\bm{p}\in\partial\Omega_{1}\setminus\mathcal{B}, the overhang angle α\alpha is defined as the angle from the base plate to the tangent line at 𝒑\bm{p} (denoted as α1\alpha_{1} and α2\alpha_{2} in Figure 1).

Refer to caption
Figure 1: Definition of overhang angle measured from the base plate to the tangent line on the boundary.

Then, there exists a critical threshold angle ψ\psi where the portion of Ω1\Omega_{1} is self-supporting (resp. requires support structures) if the overhang angle there is greater (resp. smaller) than ψ\psi. Conventional wisdom from practitioners puts the critical angle ψ\psi at 4545^{\circ} [34, 60, 82] with the reasoning that every layer then has approximately 50% contact with the previous layer and thus it is well-supported from below, although there are some authors that propose smaller values for ψ\psi (e.g., 4040^{\circ} in [61, 62]). Exact values of this critical angle also depend on the setting of the 3D printer as well as physical properties of the raw materials used. We mention that other works [67] consider an alternate definition, which is given as the angle from the vertical build direction to the outward unit normal of Ω1\partial\Omega_{1}. Denoting this as β\beta, a simple calculation shows that β=180α\beta=180^{\circ}-\alpha. Then, regions where β>180ψ\beta>180^{\circ}-\psi (typically 135135^{\circ} if ψ=45\psi=45^{\circ}) will require support structures.

In this work we adopt the latter definition, but choose to define the overhang angle as the angle measured from the negative build direction 𝒆d-\bm{e}_{d} to the outer unit normal, see Figure 2 and also [4, 30]. Then, given a critical angle threshold ψ\psi (expressed now in radians), the overhang angle constraint in the current context of shape and topology optimization would demand that an optimal geometry for the object Ω1\Omega_{1} has minimal regions where the overhang angles there do not lie in the interval [ψ,2πψ][\psi,2\pi-\psi] (i.e., Ω1\Omega_{1} should be self-supporting as much as possible), in addition to other mechanical considerations such as minimal compliance.

Refer to caption
Figure 2: The definition of overhang angle used in this work. (Left) A structure with four angles measured from the negative build direction 𝒆d-\bm{e}_{d} to the outward unit normal on the boundary. (Right) Visualization of the angles on the unit circle, where the angle ϕi\phi_{i} is associated to unit normal 𝒏i\bm{n}_{i}.

For structural topology optimization we employ the phase field methodology proposed in [28], later popularized by many authors to other applications such as multi-material structural topology optimization [22, 89], compliance optimization [24, 78], topology optimization with local stress constraints [29], nonlinear elasticity [73], elastoplasticity [6], eigenfrequency maximization [45, 78], graded-material design [32], shape optimization in fluid flow [42, 43, 44], as well as resolution strategies for some inverse identification problems [21, 57]. The key idea is to cast the structural topology optimization problem as a constrained minimization problem for a phase field variable φ\varphi, where the geometry of an optimal design Ω1\Omega_{1} for the object can be realized as a certain level-set of φ\varphi. In the language of optimal control theory, φ\varphi acts as a control and influences a response function, e.g., the elastic displacement 𝒖\bm{u} of the object, that is the solution to a system of partial differential equations (see (2.3) below). With an appropriate objective functional, for instance, a weighted sum of the mean compliance and a phase field formulation of the overhang constraint (see (2.7) below), we analyze the PDE-constrained minimization problem for an optimal design variable φ\varphi.

In the above references, previous authors employ an isotropic Ginzburg–Landau functional that has the effect of perimeter penalization (see Section 3 for more details), and in the current context this means no particular directions are preferred/discouraged in the optimization process. In this work we follow the ideas first proposed in [4, 34] that use an anisotropic perimeter functional as a geometric constraint for overhangs (see also similar ideas in [2] for support structures), and introduce a suitable anisotropic Ginzburg–Landau functional for the phase field optimization problem to enforce the overhang angle constraint. In one sense, our first novelty is that we generalize earlier works in phase field structural topology optimization by considering anisotropic energies. The precise formulation and details of the problem are given in the next sections. For more details regarding the use of anisotropy in phase field models we refer the reader to [15, 16, 46, 81].

Our second novelty is the analysis of the anisotropic phase field optimization problem. We establish the existence of minimizers (i.e., optimal designs) and derive first-order necessary optimality conditions. It turns out that our proposed approach to the overhang angle constraint requires an anisotropy function γ\gamma that is not continuously differentiable. In turn, the derivation of the associated necessary optimality conditions becomes non-standard. We overcome this difficulty with subdifferential calculus, and provide a characterization result for the subdifferential of convex functionals whose arguments are weak gradients. Then, we perform a formally matched asymptotic analysis for the optimization problem with a differentiable γ\gamma to infer the corresponding sharp interface limit.

Lastly, let us provide a non-exhaustive overview on related works that integrate the overhang angle constraint within a topology optimization procedure. Our closest counterpart is the work [4] which employs an anisotropic perimeter functional in a shape optimization framework implemented numerically with the level-set method. The authors observed that oscillations along the object boundary would develop even though the design complied with the angle constraint. This so-called dripping effect is attributed to instability effects of the anisotropic perimeter functional, brought about by the non-convexity of the anisotropy used in [4], see also Example 3.3 and Figure 10 below and the discussion in [52, Sec. 7.3]. An alternative mechanical constraint based on modeling the layer-by-layer construction process is then proposed to provide a different treatment of the overhangs and seems to suppress the dripping effect (see also [10] for similar ideas), but is computationally much more demanding. These boundary oscillations have also been observed earlier in [75], which used a Heaviside projection based integral to encode the overhang angle constraint within a density based topology optimization framework, and can be suppressed by means of an adaptive anisotropic filter introduced in [66]. Similar projection techniques for density filtering are used in [49], which was combined with an edge detection algorithm to evaluate feasible and non-feasible contours during optimization iterations. In [50], control of overhang angles is achieved by means of filtering with a wedge-shaped support, while in [92] an explicit quadratic function with respect to the design density is used to formulate the self-support constraint. A different approach was proposed in [58, 59] using a spatial filtering which checks element densities row-by-row and remove all elements not supported by the previous row. This idea is then extended in [60, 72, 83] to construct new filtering schemes that include other relevant manufacturing constraints. Another method was proposed in [51] using the frameworks of moving morphable components and moving morphable voids to provide a more explicit and geometric treatment of the problem, and is capable of simultaneously optimizing structural topology and build orientation. Finally, let us mention [85, 86], where a filter is developed based on front propagation in unstructured meshes that has a flexibility in enforcing arbitrary overhang angles. Focusing only on overhang angles in enclosed voids, [65] combined a nonlinear virtual temperature method for the identification of enclosed voids with a logarithmic-type function to constraint the area of overhang regions to zero.

The rest of this paper is organized as follows: in Section 2 we formulate the phase field structural optimization problem to be studied, and in Section 3 we describe our main idea of introducing anisotropy, along with some examples and relevant choices, as well as a useful characterization of the subdifferential of the anisotropic functional. The analysis of the structural optimization problem is carried out in Section 4, where we establish analytical results concerning minimizers and optimality conditions. The connection between our work and that of [4] is explored in Section 5 where we look into the sharp interface limit, and, finally, in Section 6 we present the numerical discretization and several simulations of our approach.

2 Problem formulation

In a bounded domain Ωd\Omega\subset\mathbb{R}^{d} with Lipschitz boundary Γ:=Ω\Gamma:=\partial\Omega that exhibits a decomposition Γ=ΓDΓgΓ0\Gamma=\Gamma_{D}\,\cup\,\Gamma_{g}\,\cup\,\Gamma_{0}, we consider a linear elastic material that does not fully occupy Ω\Omega. We describe the material location with the help of a phase field variable φ:Ω[1,1]\varphi:\Omega\to[-1,1]. In the phase field methodology, we use the level set {φ=1}={𝒙Ω:φ(𝒙)=1}\{\varphi=-1\}=\{{\bm{x}\in\Omega:\varphi(\bm{x})=-1}\} to denote the region occupied by the elastic material, and {φ=1}\{\varphi=1\} to denote the complementary void region. These two regions are separated by a diffuse interface layer {|φ|<1}\{|\varphi|<1\} whose thickness is proportional to a small parameter ε>0\varepsilon>0. Since φ\varphi describes the material distribution within Ω\Omega, complete knowledge of φ\varphi allows us to determine the shape and topology of the elastic material.

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\|\cdot\|_{L^{p}} and Wk,p\|\cdot\|_{W^{k,p}}. 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\|\cdot\|_{H^{k}}. For convenience, the norm and inner product of L2(Ω)L^{2}(\Omega) are simply denoted by \|\cdot\| and (,)(\cdot,\cdot), respectively. For our subsequent analysis, we introduce the space

HD1(Ω,d):={𝒗H1(Ω,d):𝒗=𝟎 on ΓD}.\displaystyle H^{1}_{D}(\Omega,\mathbb{R}^{d}):=\{\bm{v}\in H^{1}(\Omega,\mathbb{R}^{d})\,:\,\bm{v}=\bm{0}\text{ on }\Gamma_{D}\}.

Let us remark that, despite we employ bold symbols to denote vectors, matrices, and vector- or matrix-valued functions, we do not introduce a special notation to indicate the corresponding Lebesgue and Sobolev spaces. Thus, when those terms occur in the estimates, the corresponding norm is to be intended in its natural setting.

2.1 State system

To obtain a mathematical formulation that can be further analyzed, we employ the ersatz material approach, see [5], and model the complementary void region as a very soft elastic material. This allows us to consider a notion of elastic displacement on the entirety of Ω\Omega, leading to the displacement vector 𝒖:Ωd\bm{u}:\Omega\to\mathbb{R}^{d} and the associated linearized strain tensor

(𝒖):=12(𝒖+(𝒖)).\mathcal{E}(\bm{u}):=\tfrac{1}{2}(\nabla\bm{u}+(\nabla\bm{u})^{\top}).

Let 𝒞0\mathcal{C}_{0} and 𝒞1\mathcal{C}_{1} be the fourth order elasticity tensors corresponding to the ersatz soft material and the linear elastic material, respectively, that satisfy the standard symmetric conditions

𝒞ijkl=𝒞jikl=𝒞ijlk for i,j,k,l{1,,d},\mathcal{C}_{ijkl}=\mathcal{C}_{jikl}=\mathcal{C}_{ijlk}\quad\text{ for }i,j,k,l\in\{1,\dots,d\},

and assume there exist positive constants θ\theta and Λ\Lambda such that for any non-zero symmetric matrices 𝐀{\bf{A}}, 𝐁d×d{\bf{B}}\in\mathbb{R}^{d\times d}:

θ|𝐀|2𝒞𝐀:𝐀Λ|𝐀|2,\theta|{\bf{A}}|^{2}\leq\mathcal{C}{\bf{A}}:{\bf{A}}\leq\Lambda|{\bf{A}}|^{2},

where 𝐀:𝐁=i,j=1d𝐀ij𝐁ij{\bf{A}}:{\bf{B}}=\sum_{i,j=1}^{d}{\bf{A}}_{ij}{\bf{B}}_{ij} denotes the Frobenius inner product between two matrices. With the help of the phase field variable φ\varphi, we define an interpolation fourth order elasticity tensor 𝒞(φ)\mathcal{C}(\varphi) as

𝒞(φ)=12g(φ)(𝒞0𝒞1)+12(𝒞0+𝒞1) for φ[1,1],\mathcal{C}(\varphi)=\tfrac{1}{2}g(\varphi)(\mathcal{C}_{0}-\mathcal{C}_{1})+\tfrac{1}{2}(\mathcal{C}_{0}+\mathcal{C}_{1})\quad\text{ for }\varphi\in[-1,1],

where g:[1,1]g:\mathbb{R}\to[-1,1] is a monotone function satisfying g(1)=1g(-1)=-1 and g(1)=1g(1)=1. Then, it is clear that in the region {φ=1}\{\varphi=-1\} (resp. {φ=1}\{\varphi=1\}) we obtain the elasticity tensor 𝒞1\mathcal{C}_{1} (resp. 𝒞0\mathcal{C}_{0}) by substituting φ=1\varphi=-1 (resp. φ=1\varphi=1) into the definition of 𝒞(φ)\mathcal{C}(\varphi). As an example, we can consider

g(φ)=φ,𝒞0=ε2𝒞1,g(\varphi)=\varphi,\quad\mathcal{C}_{0}=\varepsilon^{2}\mathcal{C}_{1},

where 0<ε10<\varepsilon\ll 1 is a constant and for any symmetric matrix 𝐀d×d{\bf{A}}\in\mathbb{R}^{d\times d},

𝒞1𝐀=2μ1𝐀+λ1tr(𝐀)𝐈\displaystyle\mathcal{C}_{1}{\bf{A}}=2\mu_{1}{\bf{A}}+\lambda_{1}\mathrm{tr}({\bf{A}}){\bf I} (2.1)

with identity matrix 𝐈{\bf I} and Lamé constants λ1\lambda_{1} and μ1\mu_{1} to obtain a linear interpolating elasticity tensor. Another example for 𝒞(φ)\mathcal{C}(\varphi) uses a quadratic interpolation function

g(φ)=112(1φ)2.\displaystyle g(\varphi)=1-\frac{1}{2}(1-\varphi)^{2}. (2.2)

Let 𝒇:Ωd\bm{f}:\Omega\to\mathbb{R}^{d} denote a body force and 𝒈:Γgd\bm{g}:\Gamma_{g}\to\mathbb{R}^{d} denote a surface traction force. On ΓDΓ\Gamma_{D}\subset\Gamma we assign a zero Dirichlet boundary condition for the displacement 𝒖\bm{u} and on Γ0\Gamma_{0} we assign a traction-free boundary condition. This leads to the following elasticity system for the displacement 𝒖\bm{u}, representing the state system of the problem:

div(𝒞(φ)(𝒖))\displaystyle-\mathrm{div}(\mathcal{C}(\varphi)\mathcal{E}(\bm{u})) =𝕙(φ)𝒇\displaystyle={\mathbbm{h}}(\varphi)\bm{f}  in Ω,\displaystyle\quad\text{ in }\Omega, (2.3a)
𝒖\displaystyle\bm{u} =𝟎\displaystyle=\bm{0}  on ΓD,\displaystyle\quad\text{ on }\Gamma_{D}, (2.3b)
(𝒞(φ)(𝒖))𝒏\displaystyle(\mathcal{C}(\varphi)\mathcal{E}(\bm{u}))\bm{n} =𝒈\displaystyle=\bm{g}  on Γg,\displaystyle\quad\text{ on }\Gamma_{g}, (2.3c)
(𝒞(φ)(𝒖))𝒏\displaystyle(\mathcal{C}(\varphi)\mathcal{E}(\bm{u}))\bm{n} =𝟎\displaystyle=\bm{0}  on Γ0,\displaystyle\quad\text{ on }\Gamma_{0}, (2.3d)

where 𝒏\bm{n} denotes the outward unit normal to Γ=ΓDΓgΓ0\Gamma=\Gamma_{D}\cup\Gamma_{g}\cup\Gamma_{0}, and 𝕙:[1,1][0,1]{\mathbbm{h}}:[-1,1]\to[0,1] is defined as 𝕙(φ)=12(1φ){\mathbbm{h}}(\varphi)=\frac{1}{2}(1-\varphi) is a function introduced so that the body force 𝒇\bm{f} only acts on the elastic material {φ=1}\{\varphi=-1\} (where 𝕙(φ)=1{\mathbbm{h}}(\varphi)=1) and not on the ersatz material {φ=1}\{\varphi=1\} (where 𝕙(φ)=0{\mathbbm{h}}(\varphi)=0). We extend 𝕙(φ){\mathbbm{h}}(\varphi) as 11 if φ<1\varphi<-1 and as 0{0} if φ>1\varphi>1.

The well-posedness of (2.3) is a simple consequence of [22, Thms. 3.1, 3.2], which we summarize as follows:

Lemma 2.1.

For any φL(Ω)\varphi\in L^{\infty}(\Omega) and (𝐟,𝐠)L2(Ω,d)×L2(Γg,d)(\bm{f},\bm{g})\in L^{2}(\Omega,\mathbb{R}^{d})\times L^{2}(\Gamma_{g},\mathbb{R}^{d}), there exists a unique solution 𝐮HD1(Ω,d)\bm{u}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}) to (2.3) satisfying

Ω𝒞(φ)(𝒖):(𝒗)dx=Ω𝕙(φ)𝒇𝒗dx+Γg𝒈𝒗dd1 for all 𝒗HD1(Ω,d),\displaystyle\int_{\Omega}\mathcal{C}(\varphi)\mathcal{E}(\bm{u}):\mathcal{E}(\bm{v})\,\mathrm{dx}=\int_{\Omega}{\mathbbm{h}}(\varphi)\bm{f}\cdot\bm{v}\,\mathrm{dx}+\int_{\Gamma_{g}}\bm{g}\cdot\bm{v}\,\mathrm{d}\mathcal{H}^{d-1}\text{ for all }\bm{v}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}), (2.4)

where d1\mathcal{H}^{d-1} denotes the (d1)(d-1)-dimensional Hausdorff measure. Moreover, there exists a positive constant CC, independent of φ\varphi, such that

𝒖H1C(1+φL).\displaystyle\|\bm{u}\|_{H^{1}}\leq C(1+\|\varphi\|_{L^{\infty}}). (2.5)

In addition, let M>0M>0 and φiL(Ω)\varphi_{i}\in L^{\infty}(\Omega) with φiLM\|\varphi_{i}\|_{L^{\infty}}\leq M, i=1,2,i=1,2, and 𝐮i\bm{u}_{i} be the associated solution to (2.4). Then, there also exists a positive constant CC, depending on the data of the system and MM, but independent of the difference φ1φ2\varphi_{1}-\varphi_{2}, such that

𝒖1𝒖2H1Cφ1φ2L.\displaystyle\|\bm{u}_{1}-\bm{u}_{2}\|_{H^{1}}\leq C\|\varphi_{1}-\varphi_{2}\|_{L^{\infty}}.

The above result provides a notion of a solution operator, also referred to as the control-to-state operator, 𝑺:φ𝒖\bm{S}:\varphi\mapsto\bm{u} where 𝒖HD1(Ω,d)\bm{u}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}) is the unique solution to (2.4) corresponding to φL(Ω)\varphi\in L^{\infty}(\Omega). As such, we may view the phase field variable φ\varphi as a design variable that encodes the elastic response of the associated material distribution through the operator 𝑺\bm{S} and seek an optimal material distribution φ¯\overline{\varphi} that fulfills suitable constraints and minimizes some cost functional.

2.2 Cost functional and design space

We define the design space, i.e., the set of admissible design variables, as

𝒱m:={fH1(Ω):f[1,1] a.e. in Ω,Ωfdx=m|Ω|}H1(Ω)L(Ω),\displaystyle\mathcal{V}_{m}:=\Big{\{}f\in H^{1}(\Omega)\,:\,f\in[-1,1]\text{ a.e.~{}in }\Omega,\,\int_{\Omega}f\,\mathrm{dx}=m|\Omega|\Big{\}}\subset H^{1}(\Omega)\cap L^{\infty}(\Omega), (2.6)

where m(1,1)m\in(-1,1) is a fixed constant for the mass constraint, and motivated by the context of additive manufacturing, we propose the following cost functional to be minimized:

J(φ,𝒖)=α^Ωε2|γ(φ)|2+1εΨ(φ)dx+β(Ω𝕙(φ)𝒇𝒖dx+Γg𝒈𝒖dd1).\displaystyle J(\varphi,\bm{u})=\widehat{\alpha}\int_{\Omega}\frac{\varepsilon}{2}|\gamma(\nabla\varphi)|^{2}+\frac{1}{\varepsilon}\Psi(\varphi)\,\mathrm{dx}+\beta\Big{(}\int_{\Omega}{\mathbbm{h}}(\varphi)\bm{f}\cdot\bm{u}\,\mathrm{dx}+\int_{\Gamma_{g}}\bm{g}\cdot\bm{u}\,\mathrm{d}\mathcal{H}^{d-1}\Big{)}{.} (2.7)

In (2.7), the second term premultiplied by β>0\beta>0 is the mean compliance functional, while the first term premultiplied by α^>0\widehat{\alpha}>0 is an anisotropic Ginzburg–Landau functional with anisotropy function γ\gamma. We delay the detailed discussion on γ\gamma to the next section and remark here that for the isotropic case γ(𝒙)=|𝒙|\gamma(\bm{x})=|\bm{x}|, 𝒙d\bm{x}\in\mathbb{R}^{d}, it is well-known that the Ginzburg–Landau functional is an approximation of the perimeter functional. Therefore, (2.7) can be viewed as a weighted sum between (anisotropic) perimeter penalization and mean compliance.

Lastly, for the non-negative potential function Ψ\Psi in (2.7), we require that it has ±1\pm 1 as its global minima. While many choices are available, in light of the design space 𝒱m\mathcal{V}_{m}, we consider Ψ\Psi as the double obstacle potential

Ψ(s)={Ψ0(s):=12(1s2) if s[1,1],+ otherwise.\displaystyle\Psi(s)=\begin{cases}\Psi_{0}(s):=\frac{1}{2}(1-s^{2})&\text{ if }s\in[-1,1],\\ +\infty&\text{ otherwise}.\end{cases} (2.8)

It is worth pointing out that it holds Ψ(φ)=Ψ0(φ)\Psi(\varphi)=\Psi_{0}(\varphi) for φ𝒱m\varphi\in\mathcal{V}_{m}. Then, the structural optimization problem we study can be expressed as the following:

Minimize J(φ,𝒖)J(\varphi,\bm{u}) subject to φ𝒱m\varphi\in\mathcal{V}_{m} and 𝒖\bm{u} solving (2.3). (P)

3 Anisotropic Ginzburg–Landau functional

In the phase field methodology, the (isotropic) Ginzburg–Landau functional reads as

Eε(φ)=Ωε2|φ|2+1εΨ(φ)dx,\displaystyle E_{\varepsilon}({\varphi})=\int_{\Omega}\frac{\varepsilon}{2}|\nabla\varphi|^{2}+\frac{1}{\varepsilon}\Psi(\varphi)\,\mathrm{dx}, (3.1)

where 0<ε10<\varepsilon\ll 1 and Ψ\Psi is a non-negative double-well potential that has ±1\pm 1 as its minima, i.e., {s:Ψ(s)=0}={±1}\{s\in\mathbb{R}\,:\,\Psi(s)=0\}=\{\pm 1\}. Heuristically, the minimization of (3.1) in H1(Ω)H^{1}(\Omega) results in minimizers that take near constant values close to ±1\pm 1 in large regions of the domain Ω\Omega, which are separated by thin interfacial regions with thickness scaling with ε\varepsilon over which the functions transit smoothly from one value to the other. Formally in the limit ε0\varepsilon\to 0 these minimizers converge to functions that only take values in {±1}\{\pm 1\}. This is made rigorous in the framework of Γ\Gamma-convergence by the seminal work of Modica and Mortola [68, 69].

To facilitate the forthcoming discussion, we review some basic properties for functions of bounded variations. For a more detailed introduction we refer the reader to [8, 36]. A function uL1(Ω)u\in L^{1}(\Omega) is a function of bounded variation in Ω\Omega if its distributional gradient Du\mathrm{D}u is a finite Radon measure. The space of all such functions is denoted as BV(Ω)\mathrm{BV}(\Omega) and its endowed with the norm BV(Ω)=L1(Ω)+TV()\|\cdot\|_{\mathrm{BV}(\Omega)}=\|\cdot\|_{L^{1}(\Omega)}+\mathrm{TV}(\cdot), where for uBV(Ω)u\in\mathrm{BV}(\Omega), its total variation TV(u)\mathrm{TV}(u) is defined as

TV(u):=|Du|(Ω)=sup{Ωudivϕdx s.t. ϕC01(Ω,d),ϕL(Ω,d)1}.\mathrm{TV}(u):=|\mathrm{D}u|(\Omega)=\sup\Big{\{}\int_{\Omega}u\,\mathrm{div}{\bm{\phi}}\,\mathrm{dx}\text{ s.t. }{\bm{\phi}}\in C^{1}_{0}(\Omega,\mathbb{R}^{d}),\,\|{\bm{\phi}}\|_{{L^{\infty}(\Omega,\mathbb{R}^{d})}}\leq 1\Big{\}}.

The space BV(Ω,{a,b})\mathrm{BV}(\Omega,\{a,b\}) denotes the space of all BV(Ω)\mathrm{BV}(\Omega) functions taking values in {a,b}\{a,b\}. We say that a set UΩU\subset\Omega is a set of finite perimeter, or a Caccioppoli set, if its characteristic function χU\chi_{U}, where χU(𝒙)=1\chi_{U}(\bm{x})=1 if 𝒙U\bm{x}\in U and χU(𝒙)=0\chi_{U}(\bm{x})=0 if 𝒙U\bm{x}\notin U, belongs to BV(Ω,{0,1})\mathrm{BV}(\Omega,\{0,1\}). The perimeter of a set of finite perimeter UU in Ω\Omega is defined as

𝒫Ω(U):=|DχU|(Ω)=TV(χU),\mathcal{P}_{\Omega}(U):=|\mathrm{D}\chi_{U}|(\Omega)=\mathrm{TV}(\chi_{U}),

while its reduced boundary U\partial^{*}U is the set of all points 𝐲d{{\bf y}}\in\mathbb{R}^{d} such that |DχU|(Br(𝐲))>0|\mathrm{D}\chi_{U}|(B_{r}({{\bf y}}))>0 for all r>0r>0, with Br(𝐲)B_{r}({{\bf y}}) denoting the ball of radius rr centered at 𝐲{{\bf y}}, and

𝝂U(𝐲):=limr0DχU(Br(𝐲))|DχU|(Br(𝐲)) exists and |𝝂U(𝐲)|=1.\bm{\nu}_{U}({{\bf y}}):=\lim_{r\to 0}\frac{\mathrm{D}\chi_{U}(B_{r}({{\bf y}}))}{|\mathrm{D}\chi_{U}|(B_{r}({{\bf y}}))}\text{ exists and }|\bm{\nu}_{U}({{\bf y}})|=1.

The unit vector 𝝂U(𝐲)\bm{\nu}_{U}({{\bf y}}) is called the measure theoretical unit inner normal to UU at 𝐲{{\bf y}}, a theorem by De Giorgi yields the connection 𝒫Ω(U)=d1(U)\mathcal{P}_{\Omega}(U)=\mathcal{H}^{d-1}(\partial^{*}U), see, e.g., [8]. Then, the result of Modica and Mortola [68, 69] can be expressed as follows: The Γ\Gamma-limit of the extended functional

ε(u):={Ωε2|u|2+1εΨ(u)dx if uH1(Ω),+ elsewhere in L1(Ω),\mathcal{E}_{\varepsilon}(u):=\begin{cases}\displaystyle\int_{\Omega}\frac{\varepsilon}{2}|\nabla u|^{2}+\frac{1}{\varepsilon}\Psi(u)\,\mathrm{dx}&\text{ if }u\in H^{1}(\Omega),\\[10.0pt] +\infty&\text{ elsewhere in }L^{1}(\Omega),\end{cases}

is equal to

0(u):={cΨ𝒫Ω({u=1}) if uBV(Ω,{1,1}),+ elsewhere in L1(Ω),\mathcal{E}_{0}(u):=\begin{cases}c_{\Psi}\mathcal{P}_{\Omega}(\{u=1\})&\text{ if }u\in\mathrm{BV}(\Omega,\{-1,1\}),\\[10.0pt] +\infty&\text{ elsewhere in }L^{1}(\Omega),\end{cases}

with the constant cΨ:=112Ψ(s)𝑑sc_{\Psi}:=\int_{-1}^{1}\sqrt{2\Psi(s)}ds.

Remark 3.1.

For uBV(Ω,{1,1})u\in\mathrm{BV}(\Omega,\{-1,1\}), setting A:={u=1}A:=\{u=1\} leads to the relation u(𝐱)=2χA(𝐱)1u(\bm{x})=2\chi_{A}(\bm{x})-1, and hence

|Du|(Ω)=2|DχA|(Ω)=2𝒫Ω({u=1}).|\mathrm{D}u|(\Omega)=2|\mathrm{D}\chi_{A}|(\Omega)=2\mathcal{P}_{\Omega}(\{u=1\}).

Of particular interest is the following property of Γ\Gamma-convergence, which states that if (i) 0\mathcal{E}_{0} is the Γ\Gamma-limit of ε\mathcal{E}_{\varepsilon}, (ii) uεu_{\varepsilon} is a minimizer to ε\mathcal{E}_{\varepsilon} for every ε>0\varepsilon>0, (iii) {uε}ε>0\{u_{\varepsilon}\}_{\varepsilon>0} is a precompact sequence, then every limit of a subsequence of {uε}ε>0\{u_{\varepsilon}\}_{\varepsilon>0} is a minimizer for 0\mathcal{E}_{0}. This provides a methodology to construct minimizers of 0\mathcal{E}_{0} as limits of minimizers to ε\mathcal{E}_{\varepsilon}, provided the associated Γ\Gamma-limit is precisely 0\mathcal{E}_{0}.

Returning to our discussion and problem in additive manufacturing, in [4] it was proposed to use anisotropic perimeter functionals to model the overhang angle constraint. In our notation, for a set UU of bounded variation with 𝝂U\bm{\nu}_{U} as the measure theoretical inward unit normal, these functionals take the form

𝒫γ(U):=Uγ(𝝂U)dd1\displaystyle\mathcal{P}_{\gamma}(U):=\int_{\partial^{*}U}\gamma(\bm{\nu}_{U})\,\mathrm{d}\mathcal{H}^{d-1} (3.2)

with a C1C^{1} function γ:d\gamma:\mathbb{R}^{d}\to\mathbb{R}. Two choices were suggested in [4]:

γa(𝝂):=[min(0,𝝂𝒆d+cosψ)]2,γb(𝝂):=i=1m(𝝂𝝂ψi)2,\displaystyle\gamma_{a}(\bm{\nu}):=[\min(0,\bm{\nu}\cdot\bm{e}_{d}+\cos\psi)]^{2},\quad\gamma_{b}(\bm{\nu}):=\prod_{i=1}^{m}(\bm{\nu}-\bm{\nu}_{\psi_{i}})^{2},

where ψ\psi is a fixed angle threshold, 𝒆d\bm{e}_{d} denotes the build direction, and ψi:d\psi_{i}:\mathbb{R}^{d}\to\mathbb{R}, for i=1,,mi=1,\dots,m, are given pattern functions with 𝝂ψi:=ψi/|ψi|\bm{\nu}_{\psi_{i}}:=\nabla\psi_{i}/|\nabla\psi_{i}|. The first choice γa\gamma_{a} penalizes the regions of the boundary U\partial^{*}U where the angle between the outward normal (𝝂U)(-\bm{\nu}_{U}) and the negative build direction (𝒆d)(-\bm{e}_{d}) is smaller than ψ\psi, while the second choice γb\gamma_{b} compels the unit normal 𝝂U\bm{\nu}_{U} to be close to at least one of the directions 𝝂ψi\bm{\nu}_{\psi_{i}}.

A phase field approximation of the anisotropic perimeter functional (3.2) is the following anisotropic Ginzburg–Landau functional

Eγ,ε(φ):=Ωε2|γ(φ)|2+1εΨ(φ)dx,\displaystyle E_{\gamma,\varepsilon}(\varphi):=\int_{\Omega}\frac{\varepsilon}{2}|\gamma(\nabla\varphi)|^{2}+\frac{1}{\varepsilon}\Psi(\varphi)\,\mathrm{dx}, (3.3)

where as before, Ψ\Psi is a double well potential with ±1\pm 1 as its minima. Then, for a convex γ:d\gamma:\mathbb{R}^{d}\to\mathbb{R} that is positively homogeneous of degree one (see (A1)), one has the analogue of the result by Modica and Mortola for anisotropic energies (see [19, 20, 27, 71], and also Lemma 5.1 below), that is

Γlimε0γ,ε(u)=γ,0(u),\displaystyle\Gamma-\lim_{\varepsilon\to 0}\mathcal{E}_{\gamma,\varepsilon}(u)=\mathcal{E}_{\gamma,0}(u),

where the extended functionals γ,ε\mathcal{E}_{\gamma,\varepsilon} and γ,0\mathcal{E}_{\gamma,0} are defined as

γ,ε(u)\displaystyle\mathcal{E}_{\gamma,\varepsilon}(u) :={Ωε2|γ(u)|2+1εΨ(u)dx if uH1(Ω),+ elsewhere in L1(Ω),\displaystyle:=\begin{cases}\displaystyle\int_{\Omega}\frac{\varepsilon}{2}|\gamma(\nabla u)|^{2}+\frac{1}{\varepsilon}\Psi(u)\,\mathrm{dx}&\text{ if }u\in H^{1}(\Omega),\\[10.0pt] +\infty&\text{ elsewhere in }L^{1}(\Omega),\end{cases} (3.4a)
γ,0(u)\displaystyle\mathcal{E}_{\gamma,0}(u) :={cΨPγ({u=1}) if uBV(Ω,{1,1}),+ elsewhere in L1(Ω).\displaystyle:=\begin{cases}c_{\Psi}P_{\gamma}(\{u=1\})&\text{ if }u\in\mathrm{BV}(\Omega,\{-1,1\}),\\[10.0pt] +\infty&\text{ elsewhere in }L^{1}(\Omega).\end{cases} (3.4b)

This motivates our consideration of the objective functional (2.7) and of the study of the related minimization problem for the overhang angle constraint. Furthermore, let us formally state the corresponding sharp interface limit (ε0)(\varepsilon\to 0) of the structural optimization problem as:

Minimize J0(φ,𝒖)J_{0}(\varphi,\bm{u}) subject to φBVm(Ω,{1,1})\varphi\in\mathrm{BV}_{m}(\Omega,\{-1,1\}) and 𝒖\bm{u} solving (2.3), (𝐏0\mathbf{P}_{0})

where BVm(Ω,{1,1})={fBV(Ω,{1,1}):Ωfdx=m|Ω|}\mathrm{BV}_{m}(\Omega,\{-1,1\})=\{f\in\mathrm{BV}(\Omega,\{-1,1\})\,:\,\int_{\Omega}f\,\mathrm{dx}=m|\Omega|\} and

J0(φ,𝒖)=α^cΨ𝒫γ({φ=1})+β(Ω𝕙(φ)𝒇𝒖dx+Γg𝒈𝒖dd1).J_{0}(\varphi,\bm{u})={\widehat{\alpha}}c_{\Psi}{\cal P}_{\gamma}(\{\varphi=1\})+\beta\Big{(}\int_{\Omega}{\mathbbm{h}}(\varphi)\bm{f}\cdot\bm{u}\,\mathrm{dx}+\int_{\Gamma_{g}}\bm{g}\cdot\bm{u}\,\mathrm{d}\mathcal{H}^{d-1}\Big{)}.

Note that by Lemma 2.1, the solution operator 𝑺:φ𝒖\bm{S}:\varphi\mapsto\bm{u} is well-defined for φBV(Ω,{1,1})\varphi\in\mathrm{BV}(\Omega,\{-1,1\}). The connection between (P) and (𝐏0\mathbf{P}_{0}) will be explored in Section 5.

3.1 Anisotropy function, Wulff shape and Frank diagram

Consider an anisotropic density function γ:d\gamma:\mathbb{R}^{d}\to\mathbb{R} satisfying

  1. (A1)

    γ\gamma is positively homogeneous of degree one:

    γ(λ𝒒)=λγ(𝒒) for all 𝒒d{𝟎},λ0,\displaystyle\gamma(\lambda\bm{q})=\lambda\gamma(\bm{q})\quad\text{ for all }\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\},\,\lambda\geq 0,

    which immediately implies γ(𝟎)=0\gamma(\bm{0})=0.

  2. (A2)

    γ\gamma is positive for non-zero vectors:

    γ(𝒒)>0 for all 𝒒d{𝟎}.\displaystyle\gamma(\bm{q})>0\quad\text{ for all }\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\}.
  3. (A3)

    γ\gamma is convex:

    γ(s𝒑+(1s)𝒒)sγ(𝒑)+(1s)γ(𝒒) for all 𝒑,𝒒d,s[0,1].\gamma(s\bm{p}+(1-s)\bm{q})\leq s\gamma(\bm{p})+(1-s)\gamma(\bm{q})\quad\text{ for all }\bm{p},\bm{q}\in\mathbb{R}^{d},\,s\in[0,1].

Note that it is sufficient to assign values of γ\gamma on the unit sphere B1(𝟎)\partial B_{1}(\bm{0}) in d\mathbb{R}^{d}, since by the one-homogeneity property (A1) we can define for any 𝒑𝟎\bm{p}\neq\bm{0} with 𝒑^=𝒑/|𝒑|\hat{\bm{p}}=\bm{p}/|\bm{p}|

γ(𝒑):=|𝒑|γ(𝒑^).\gamma(\bm{p}):=|\bm{p}|\gamma\big{(}\hat{\bm{p}}\big{)}.

A consequence of the convexity assumption (A3) is that γ\gamma is continuous (in fact it is even locally Lipschitz continuous, see, e.g., [7, E4.6, p. 129]). Then, from (A2) we have that γ\gamma has a positive minimum c~\tilde{c} on the compact set B1(𝟎)\partial B_{1}(\bm{0}), and consequently by (A1), γ(𝒒)=γ(|𝒒|𝒒^)=|𝒒|γ(𝒒^)c~|𝒒|\gamma(\bm{q})=\gamma\big{(}|\bm{q}|\hat{\bm{q}}\big{)}=|\bm{q}|\gamma\big{(}\hat{\bm{q}}\big{)}\geq\tilde{c}|\bm{q}| for 𝒒𝟎\bm{q}\neq\bm{0}. Thus, (A1)–(A3) yield the following property:

  1. (A4)

    γ\gamma is Lipschitz continuous and there exists a constant c~>0\tilde{c}>0 such that

    γ(𝒒)c~|𝒒| for all 𝒒d.\displaystyle\gamma(\bm{q})\geq\tilde{c}|\bm{q}|\quad\text{ for all }\bm{q}\in\mathbb{R}^{d}.

Moreover, with λ=2\lambda=2 in (A1) and s=12s=\frac{1}{2} in (A3), it is not difficult to verify that γ\gamma satisfies the triangle inequality. Hence, provided that γ(𝒒)=γ(𝒒)\gamma(\bm{q})=\gamma(-\bm{q}) for all 𝒒d\bm{q}\in\mathbb{R}^{d}, any γ\gamma satisfying (A1)–(A3) defines a norm on d\mathbb{R}^{d}.

For such anisotropy density functions and smooth hypersurfaces Γ\Gamma with normal vector field 𝝂\bm{\nu}, we define the anisotropic interfacial energy as

γ(Γ):=Γγ(𝝂)dd1.\displaystyle\mathcal{F}^{\gamma}(\Gamma):=\int_{\Gamma}\gamma(\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1}. (3.5)

Then, the isoperimetric problem involves finding a hypersurface Γ\Gamma^{*} that minimizes γ\mathcal{F}^{\gamma} under a volume constraint. This problem has been well-studied by many authors (see for instance [39, 40, 79, 80] and [52] and the references therein) and the solution is given as the boundary of the region called the Wulff shape [90]

W={𝒓d:γ(𝒓)1},\displaystyle W=\{\bm{r}\in\mathbb{R}^{d}\,:\,\gamma^{*}(\bm{r})\leq 1\}, (3.6)

where γ\gamma^{*} is the dual function of γ\gamma defined as

γ(𝒓)=sup𝒒d{𝟎}𝒓𝒒γ(𝒒) for all 𝒓d.\displaystyle\gamma^{*}(\bm{r})=\sup_{\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\}}\frac{\bm{r}\cdot\bm{q}}{\gamma(\bm{q})}\quad\text{ for all }\bm{r}\in\mathbb{R}^{d}. (3.7)

The dual function γ\gamma^{*} also satisfies (A1)–(A3) and hence we can view the Wulff shape WW as the 1-ball of γ\gamma^{*}. Besides the Wulff shape, another region of interest that is used to visualize the effects of the anisotropy is the Frank diagram [38], which is defined as the 1-ball of γ\gamma:

F={𝒒d:γ(𝒒)1},\displaystyle F=\{\bm{q}\in\mathbb{R}^{d}\,:\,\gamma(\bm{q})\leq 1\}, (3.8)

which, due to (A3), is always a convex subset of d\mathbb{R}^{d}. To see how the shape of the boundary of FF determines which directions of unit sphere in d\mathbb{R}^{d} is preferred by the anisotropic density function γ\gamma, let us consider three examples.

Example 3.1 (Isotropic case).

Consider γ(𝐪)=|𝐪|\gamma(\bm{q})=|\bm{q}| for 𝐪d\bm{q}\in\mathbb{R}^{d}. Then, the associated Frank diagram is just the unit ball in d\mathbb{R}^{d}, with boundary {γ(𝐪)=|𝐪|=1}\{\gamma(\bm{q})=|\bm{q}|=1\}. As all points on the boundary are equidistant to the origin, all directions of the unit sphere in d\mathbb{R}^{d} are equally preferable.

Example 3.2 (Convex example).

Consider the Frank diagram shown in the left of Figure 3, whose boundary is composed of a circular arc CC and a horizontal line LL. The black dot denotes the origin in 2\mathbb{R}^{2}. For any unit vector in 2\mathbb{R}^{2}, we denote by ϕ[0,2π)\phi\in[0,2\pi) the angle it makes with the negative yy-axis measured anticlockwise (see also Figure 2). Then, there exists θ>0\theta>0 such that all unit vectors with angle in [0,θ][2πθ,2π)[0,\theta]\cup[2\pi-\theta,2\pi) are associated with the horizontal line LL in Figure 3, while all unit vectors with angle in (θ,2πθ)(\theta,2\pi-\theta) are associated with the circular arc CC. Notice, if the origin lies on LL, then θ=π2\theta=\frac{\pi}{2}, and CC is the upper semicircle.

Let 𝐩C\bm{p}\in C and 𝐪L\bm{q}\in L be arbitrary, and set 𝐩^=𝐩|𝐩|\hat{\bm{p}}=\frac{\bm{p}}{|\bm{p}|}, 𝐪^=𝐪|𝐪|\hat{\bm{q}}=\frac{\bm{q}}{|\bm{q}|} as their unit vectors. It is clear from the figure that |𝐩||𝐪||\bm{p}|\geq|\bm{q}|, and since γ(𝐩)=γ(𝐪)=1\gamma(\bm{p})=\gamma(\bm{q})=1, by (A1) we see that

1=|𝒑|γ(𝒑^)=|𝒒|γ(𝒒^).\displaystyle 1=|\bm{p}|\gamma\big{(}\hat{\bm{p}}\big{)}=|\bm{q}|\gamma\big{(}\hat{\bm{q}}\big{)}.

This implies that γ(𝐩^)γ(𝐪^)\gamma(\hat{\bm{p}})\leq\gamma(\hat{\bm{q}}), and from the viewpoint of minimizing the interfacial energy γ\mathcal{F}^{\gamma} in (3.5), directions 𝐩^\hat{\bm{p}} are preferable to directions 𝐪^\hat{\bm{q}}. Consequently, from the Frank diagram in Figure 3 we can see that the associated anisotropy density function γ\gamma prefers directions with angle in (θ,2πθ)(\theta,2\pi-\theta) over directions with angle in [0,θ][2πθ,2π)[0,\theta]\cup[2\pi-\theta,2\pi).

Refer to caption
Figure 3: Illustrations of the Frank diagram for Examples 3.2 and 3.3. (Left) Convex case and (Right) non-convex case.
Example 3.3 (Non-convex example).

Consider the Frank diagram shown in the right of Figure 3, whose boundary encloses a non-convex set. Let 𝐫1\bm{r}_{1} and 𝐫2\bm{r}_{2} denote the two unit vectors whose angles, say θ\theta and 2πθ2\pi-\theta, respectively, associate to the two endpoints of the circular arc. From previous discussions, the anisotropy density function γ\gamma will prefer directions with angles in (θ,2πθ)(\theta,2\pi-\theta).

With such γ\gamma, consider two spatial points 𝐱1\bm{x}_{1} and 𝐱2\bm{x}_{2} at the same height, see Figure 4. Connecting them via a horizontal straight line is energetically expensive since this is associated to a direction with angle zero (where on the boundary of the Frank diagram is closest to the origin). An energetically more favorable connection is a zigzag path from 𝐱1\bm{x}_{1} and 𝐱2\bm{x}_{2} whose normal vectors oscillate between 𝐫1\bm{r}_{1} and 𝐫2\bm{r}_{2}. This is similar to a behavior termed “dripping effect” in [4] (see also [75, Fig. 15]), which is the tendency for shapes to develop oscillatory boundaries in order to meet the overhang angle constraints.

Refer to caption
Figure 4: Connection between two spatial points with non-convex anisotropic density function γ\gamma whose Frank diagram looks like the right of Figure 3.

3.2 Relevant examples of anisotropic density function

Motivated by the above discussion, in this section we provide some examples of γ\gamma that achieve the Frank diagrams shown in Figure 3.

3.2.1 An example of a convex anisotropy

To fix the ideas, let us begin with the two-dimensional case. For a fixed constant α(0,1)\alpha\in(0,1), we consider the function

γα(𝒙)={x12+x22 if 𝒙=(x1,x2)V,1αx2 if (x1,x2)2V¯=:Vc,\displaystyle\gamma_{\alpha}(\bm{x})=\begin{cases}\sqrt{x_{1}^{2}+x_{2}^{2}}&\text{ if }\bm{x}=(x_{1},x_{2})\in V,\\[10.0pt] \displaystyle-\frac{1}{\alpha}x_{2}&\text{ if }(x_{1},x_{2})\in\mathbb{R}^{2}\setminus\overline{V}=:V^{c},\end{cases} (3.9)

where the set V2V{\subseteq\mathbb{R}^{2}} will be determined in the following. To achieve the boundary of the Frank diagram FF shown in the left of Figure 3, we notice that FV\partial F\cap V is a circular arc of radius 11 centered at the origin, while FVc\partial F\cap V^{c} implies x2=αx_{2}=-\alpha which is a horizontal line segment at height α-\alpha. Following the convention in Figure 2 where angles are measured anticlockwise from the negative x2x_{2}-axis, we can parameterize the circular arc by (cosϕ,sinϕ)(\cos\phi,\sin\phi) for ϕ[θ,2πθ]\phi\in[\theta,2\pi-\theta] where θ:=cos1(α)\theta:=\cos^{-1}(\alpha), see Figure 3. Hence, the set VV in the definition (3.9) can be characterized as

V={cosϕα:ϕ[0,2π]}.\displaystyle V=\Big{\{}\cos\phi\leq\alpha\,:\,\phi\in[0,2\pi]\Big{\}}.
Remark 3.2.

Note that in the limiting case α=1\alpha=1, the above set VV is the entire plane 2\mathbb{R}^{2}. Hence we have the isotropic case γ(𝐱)=|𝐱|\gamma(\bm{x})=|\bm{x}| for all 𝐱2\bm{x}\in\mathbb{R}^{2}, and F\partial F is simply the unit circle.

For a parametric characterization of the set VV, let c:=α1α2c:=\frac{\alpha}{\sqrt{1-\alpha^{2}}} and consider the two straight lines {x2=cx1}\{x_{2}=cx_{1}\} and {x2=cx1}\{x_{2}=-cx_{1}\} dividing 2\mathbb{R}^{2} into eight regions (see Figure 5), which we label as Region 1,2,,81,2,\dots,8 in an anticlockwise direction starting from the positive x1x_{1}-axis. Then, the horizontal straight line portion FVc\partial F\cap V^{c} of the Frank diagram at height x2=αx_{2}=-\alpha is contained in Regions 6 and 7, whose union is described by the set {x2<c|x1|}\{x_{2}<-c|x_{1}|\}, while the circular arc portion FV\partial F\cap V is contained in Regions 1,,51,\dots,5 and 88, whose union is described by the set {x2c|x1|}\{x_{2}\geq-c|x_{1}|\}. Hence, a parameteric characterization of the set VV in (3.9) is

V={x2α1α2|x1|}={x2α|𝒙|}.\displaystyle V=\Big{\{}x_{2}\geq-\frac{\alpha}{\sqrt{1-\alpha^{2}}}|x_{1}|\Big{\}}=\Big{\{}x_{2}\geq-\alpha|\bm{x}|\Big{\}}.

Generalizing to the dd-dimensional case, we obtain

γα(𝒙)={|𝒙| if xdα|𝒙|1αxd if xd<α|𝒙| for 𝒙=(x1,,xd)d.\displaystyle\gamma_{\alpha}(\bm{x})=\begin{cases}|\bm{x}|&\text{ if }\displaystyle x_{d}\geq-\alpha|\bm{x}|\\[10.0pt] \displaystyle-\frac{1}{\alpha}x_{d}&\text{ if }\displaystyle x_{d}<-\alpha|\bm{x}|\end{cases}\text{ for }\bm{x}=(x_{1},\dots,x_{d})\in\mathbb{R}^{d}. (3.10)

From (3.10) we see that γαC0(d)\gamma_{\alpha}\in C^{0}(\mathbb{R}^{d}) but it is not continuously differentiable at the points xd=α|𝒙|x_{d}=-\alpha|\bm{x}|. Furthermore, it is clear that γα\gamma_{\alpha} satisfies the assumptions (A1)–(A3) and hence (A4).

Refer to caption
Figure 5: Schematics for the parametric characterization.

3.2.2 An example of a non-convex anisotropy

For completeness, we provide an example of an anisotropic function γ\gamma that yields a non-convex Frank diagram as seen in the right of Figure 3. Referring to the construction of the previous example, we present the two-dimensional case first. Fix λ(0,1)\lambda\in(0,1) and let (0,λα)(0,-\lambda\alpha)^{\top} denote the intersection of the two straight line segments (which we call L1L_{1} and L2L_{2} respectively) just below the origin. Denoting by 𝒑\bm{p}_{-} the endpoint of the left line segment L1L_{1} and by 𝒑+\bm{p}_{+} the endpoint of the right line segment L2L_{2} that connects (0,λα)(0,-\lambda\alpha)^{\top} to the circular arc of radius 11, a short calculation shows that 𝒑±=(±1α2,α)\bm{p}_{\pm}=(\pm\sqrt{1-\alpha^{2}},-\alpha)^{\top}.

A choice of tangent vector for L1L_{1} is 𝝉=(1α2,(λ1)α)\bm{\tau}=(-\sqrt{1-\alpha^{2}},(\lambda-1)\alpha)^{\top} so that a normal vector for L1L_{1} is 𝒏=((1λ)α,1α2)\bm{n}=((1-\lambda)\alpha,-\sqrt{1-\alpha^{2}})^{\top}. Consider, for some constant b>0b>0 to be identified,

γ(𝒙)=b|(1λ)αx1x21α2|\gamma(\bm{x})=b\big{|}(1-\lambda)\alpha x_{1}-x_{2}\sqrt{1-\alpha^{2}}\big{|}

for 𝒙=(x1,x2){x10,x2α1α2x1}\bm{x}=(x_{1},x_{2})\in\{x_{1}\leq 0,x_{2}\leq\frac{\alpha}{\sqrt{1-\alpha^{2}}}x_{1}\}. This corresponds to Region 6 in the right of Figure 5 that contains the line segment L1L_{1}, which can be parameterized as

L1={(ζ1α2,λαζ(1λ)α):ζ[0,1]}.\displaystyle L_{1}=\big{\{}(-\zeta\sqrt{1-\alpha^{2}},-\lambda\alpha-\zeta(1-\lambda)\alpha)^{\top}\,:\,\zeta\in[0,1]\big{\}}.

Notice that γ\gamma satisfies (A1)–(A2) (due to the modulus), and a short calculation shows that for 𝒙L1\bm{x}\in L_{1},

γ(𝒙)=b|λα1α2|=bλα1α2.\gamma(\bm{x})=b\big{|}\lambda\alpha\sqrt{1-\alpha^{2}}\big{|}=b\lambda\alpha\sqrt{1-\alpha^{2}}.

Hence, choosing b=(λα1α2)1>0b=(\lambda\alpha\sqrt{1-\alpha^{2}})^{-1}>0 yields that γ(𝒙)=1\gamma(\bm{x})=1 for 𝒙L1\bm{x}\in L_{1}. Similarly, a choice of tangent vector for L2L_{2} is 𝝉=(1α2,(λ1)α)\bm{\tau}=(\sqrt{1-\alpha^{2}},(\lambda-1)\alpha)^{\top}, so that a normal vector for L2L_{2} is 𝒏=((λ1)α,1α2)\bm{n}=((\lambda-1)\alpha,-\sqrt{1-\alpha^{2}})^{\top}. We consider

γ(𝒙)\displaystyle\gamma(\bm{x}) =1λα1α2|(λ1)αx1x21α2|\displaystyle=\frac{1}{\lambda\alpha\sqrt{1-\alpha^{2}}}\big{|}(\lambda-1)\alpha x_{1}-x_{2}\sqrt{1-\alpha^{2}}\big{|}
=1λα1α2|(1λ)αx1+x21α2|\displaystyle=\frac{1}{\lambda\alpha\sqrt{1-\alpha^{2}}}\big{|}(1-\lambda)\alpha x_{1}+x_{2}\sqrt{1-\alpha^{2}}\big{|}

for 𝒙{x10,x2α1α2x1}\bm{x}\in\{x_{1}\geq 0,x_{2}\leq-\frac{\alpha}{\sqrt{1-\alpha^{2}}}x_{1}\} which corresponds to Region 7 in Figure 5 that contains the line segment L2L_{2}. Then, a short calculation shows that γ(𝒙)=1\gamma(\bm{x})=1 for 𝒙L2\bm{x}\in L_{2}. Thus, an example of an anisotropic function γ\gamma that give rise to a Frank diagram whose boundary is the right figure in Figure 3 is

γα,λ(𝒙)={|𝒙| if x2α|𝒙|,|1λλ1α2x11λαx2| if x10,x2<α|𝒙|,|1λλ1α2x1+1λαx2| if x1>0,x2<α|𝒙|,\displaystyle\gamma_{\alpha,\lambda}(\bm{x})=\begin{cases}|\bm{x}|&\text{ if }\displaystyle x_{2}\geq-\alpha|\bm{x}|,\\[10.0pt] \displaystyle\Big{|}\frac{1-\lambda}{\lambda\sqrt{1-\alpha^{2}}}x_{1}-\frac{1}{\lambda\alpha}x_{2}\Big{|}&\text{ if }\displaystyle x_{1}\leq 0,\,x_{2}<-\alpha|\bm{x}|,\\[10.0pt] \displaystyle\Big{|}\frac{1-\lambda}{\lambda\sqrt{1-\alpha^{2}}}x_{1}+\frac{1}{\lambda\alpha}x_{2}\Big{|}&\text{ if }\displaystyle x_{1}>0,\,x_{2}<-\alpha|\bm{x}|,\end{cases} (3.11)

for α(0,1)\alpha\in(0,1), λ(0,1)\lambda\in(0,1) and 𝒙=(x1,x2)2\bm{x}=(x_{1},x_{2})\in\mathbb{R}^{2}. Notice that in the limit λ1\lambda\to 1, we recover the convex anisotropic function γα\gamma_{\alpha} defined in (3.10).

To generalize to the dd-dimensional case, we notice that the lines L1L_{1} and L2L_{2} in the above discussion are now replaced by the lateral surface SS of a cone with apex (0,0,,0,λα)d(0,0,\dots,0,-\lambda\alpha)\in\mathbb{R}^{d}, which can be parameterized as

S={𝒙=(𝒙~,xd)d:xd+λα=(1λ)αR2α2|𝒙~|,xd[α,λα]}.S=\left\{\bm{x}=(\tilde{\bm{x}},x_{d})\in\mathbb{R}^{d}\,:\,x_{d}+\lambda\alpha=-\frac{(1-\lambda)\alpha}{\sqrt{R^{2}-\alpha^{2}}}|\tilde{\bm{x}}|,\,x_{d}\in[-\alpha,-\lambda\alpha]\right\}.

Then, by similar arguments leading to (3.11), we obtain the function

γα,λ(𝒙)={|𝒙| if xdα|𝒙|,|1λλR2α2|𝒙~|+1λαxd| if xd<α|𝒙|,\displaystyle\gamma_{\alpha,\lambda}(\bm{x})=\begin{cases}\displaystyle|\bm{x}|&\text{ if }\displaystyle x_{d}\geq-\alpha|\bm{x}|,\\[10.0pt] \displaystyle\Big{|}\frac{1-\lambda}{\lambda\sqrt{R^{2}-\alpha^{2}}}|\tilde{\bm{x}}|+\frac{1}{\lambda\alpha}x_{d}\Big{|}&\text{ if }\displaystyle x_{d}<-\alpha|\bm{x}|,\end{cases}

for 𝒙=(𝒙~,xd)d\bm{x}=(\tilde{\bm{x}},x_{d})\in\mathbb{R}^{d}, 𝒙~d1\tilde{\bm{x}}\in\mathbb{R}^{d-1}, where we can verify that γα,λ(𝒙)=1\gamma_{\alpha,\lambda}(\bm{x})=1 for 𝒙S\bm{x}\in S. In Figure 6 we display the Frank diagrams for non-convex anisotropy functions of the form (3.11) with λ=0.5\lambda=0.5 and α=0.7,0.5,0.3\alpha{=0.7,0.5,0.3}.

Refer to caption
Refer to caption
Refer to caption
Figure 6: Frank diagrams for non-convex anisotropy (3.11) with λ=0.5\lambda=0.5. From left to right: α=0.7\alpha=0.7, 0.50.5 and 0.30.3.

3.3 Subdifferential characterization

The analysis of the structural optimization problem (P) follows almost analogously as in [22]. The major difference is the anisotropic Ginzburg–Landau functional in (2.7). From the examples of γ\gamma discussed in the previous subsection, our interest lies in anisotropy density functions that are convex and continuous, but not necessarily C1(d)C^{1}(\mathbb{R}^{d}), which necessitates a non-trivial modification to the analysis performed in [22]. In this subsection we focus only on the gradient part of (2.7) and investigate its subdifferential in preparation for the first-order necessary optimality conditions for (P) (cf. Theorem 4.2).

We define A:d[0,)A:\mathbb{R}^{d}\to[0,\infty) as

A(𝒒)=12|γ(𝒒)|2 for all 𝒒d,\displaystyle A(\bm{q})=\frac{1}{2}|\gamma(\bm{q})|^{2}\quad\text{ for all }\bm{q}\in\mathbb{R}^{d}, (3.12)

and from (A1)–(A4), it readily follows that AA is convex, continuous with A(𝟎)=0A(\bm{0})=0, positive for non-zero vectors, positively homogeneous of degree two:

A(λ𝒒)=λ2A(𝒒) for all 𝒒d{𝟎},λ0,A(\lambda\bm{q})=\lambda^{2}A(\bm{q})\quad\text{ for all }\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\},\quad\lambda\geq 0,

and there exist positive constants cc and CC such that

A(𝒒)c|𝒒|2,|A(𝒒)A(𝒓)|C|𝒒𝒓|(|𝒒|+|𝒓|) for all 𝒒,𝒓d.\displaystyle A(\bm{q})\geq c|\bm{q}|^{2},\quad|A(\bm{q})-A(\bm{r})|\leq C|\bm{q}-\bm{r}|(|\bm{q}|+|\bm{r}|)\quad\text{ for all }\bm{q},\bm{r}\in\mathbb{R}^{d}. (3.13)

Associated to such a function AA, we consider the integral functional

:H1(Ω),φΩA(φ)dx.\displaystyle\mathcal{F}:H^{1}(\Omega)\to\mathbb{R},\quad\varphi\mapsto\int_{\Omega}A(\nabla\varphi)\,\mathrm{dx}. (3.14)

Convexity of AA immediately imply that \mathcal{F} is convex, proper and weakly lower semicontinuous in H1(Ω)H^{1}(\Omega), see, e.g., [35, Thm. 1, §8.2.2]. Consequently, its subdifferential :H1(Ω)2H1(Ω)\partial\mathcal{F}:H^{1}(\Omega)\to 2^{H^{1}(\Omega)^{*}}, defined as

(φ)={ξH1(Ω):(ϕ)(φ)ξ,ϕφH1ϕH1(Ω)},\displaystyle\partial\mathcal{F}(\varphi)=\Big{\{}\xi\in H^{1}(\Omega)^{*}\,:\,\mathcal{F}(\phi)-\mathcal{F}(\varphi)\geq\langle\xi,\phi-\varphi\rangle_{H^{1}}\,\;\forall\phi\in H^{1}(\Omega)\Big{\}},

is a maximal monotone operator from H1(Ω)H^{1}(\Omega) to H1(Ω)H^{1}(\Omega)^{*} (see [13, Thm. 2.43]), which is equivalent to the property that, for any fH1(Ω)f\in H^{1}(\Omega)^{*}, there exists at least one solution φ0D()\varphi_{0}\in D(\partial\mathcal{F}) with ξ0(φ0)\xi_{0}\in\partial\mathcal{F}(\varphi_{0}) such that

φ0+ξ0=f in H1(Ω).\displaystyle\varphi_{0}+\xi_{0}=f\quad\text{ in }H^{1}(\Omega)^{*}.

Let us now provide a useful lemma that characterizes the subdifferential (φ)\partial\mathcal{F}(\varphi) in terms of elements of A(φ)\partial A(\nabla\varphi) (see also [13, p. 146, Problem 2.7]). Heuristically, elements of (φ)\partial\mathcal{F}(\varphi) are the negative weak divergences of elements of A(φ)\partial A(\nabla\varphi).

Lemma 3.1.

Consider the map 𝒢:H1(Ω)2H1(Ω)\mathcal{G}:H^{1}(\Omega)\to 2^{H^{1}(\Omega)^{*}} defined by

φ{vH1(Ω):\displaystyle\varphi\mapsto\Big{\{}v\in H^{1}(\Omega)^{*}\,:\, 𝝃L2(Ω,d) s.t. 𝝃A(φ) a.e. in Ω,\displaystyle\exists\,\bm{\xi}\in L^{2}(\Omega,\mathbb{R}^{d})\text{ s.t. }\bm{\xi}\in\partial A(\nabla\varphi)\text{ a.e. in }\Omega,
and v,pH1=(𝝃,p)pH1(Ω)}.\displaystyle\text{ and }\langle v,p\rangle_{H^{1}}=(\bm{\xi},\nabla p)\;\forall p\in H^{1}(\Omega)\Big{\}}.

Then, for φD()\varphi\in D(\partial\mathcal{F}), it holds that

(φ)=𝒢(φ).\displaystyle\partial\mathcal{F}(\varphi)=\mathcal{G}(\varphi).
Remark 3.3.

If A:d[0,)A:\mathbb{R}^{d}\to[0,\infty) is smooth, convex, coercive with bounded second derivatives, i.e., there exists C>0C>0 such that |Api,pj(𝐩)|C|A_{p_{i},p_{j}}(\bm{p})|\leq C for all 𝐩d\bm{p}\in\mathbb{R}^{d} and 1i,jd1\leq i,j\leq d, then the corresponding characterization of the subdifferential of \mathcal{F} can be found in [35, Thm. 4, §9.6.3]. Namely, (φ)\partial\mathcal{F}(\varphi) is single-valued and (φ)={i=1dxi[Api(φ)]}\partial\mathcal{F}(\varphi)=\{-\sum_{i=1}^{d}\partial_{x_{i}}[A_{p_{i}}(\nabla\varphi)]\}.

Proof of Lemma 3.1.

To begin, fix φH1(Ω)\varphi\in H^{1}(\Omega), v𝒢(φ)v\in\mathcal{G}(\varphi) and let 𝝃L2(Ω,d)\bm{\xi}\in L^{2}(\Omega,\mathbb{R}^{d}) with 𝝃(𝒙)A(φ(𝒙))\bm{\xi}(\bm{x})\in\partial A(\nabla\varphi(\bm{x})) for almost every 𝒙Ω\bm{x}\in\Omega satisfying v,pH1=(𝝃,p)\langle v,p\rangle_{H^{1}}=(\bm{\xi},\nabla p) for all pH1(Ω)p\in H^{1}(\Omega). By the definition of the subgradient A(φ(𝒙))\partial A(\nabla\varphi(\bm{x})), we have

A(𝒒)A(φ(𝒙))𝝃(𝒙)(𝒒φ(𝒙))𝒒d.\displaystyle A(\bm{q})-A(\nabla\varphi(\bm{x}))\geq\bm{\xi}(\bm{x})\cdot(\bm{q}-\nabla\varphi(\bm{x}))\quad\forall\bm{q}\in\mathbb{R}^{d}.

Choosing 𝒒=ϕ(𝒙)\bm{q}=\nabla\phi(\bm{x}) where ϕH1(Ω)\phi\in H^{1}(\Omega) is arbitrary, and integrating over Ω\Omega we infer that

(ϕ)(φ)Ω𝝃(ϕφ)dx=v,ϕφH1ϕH1(Ω).\displaystyle\mathcal{F}(\phi)-\mathcal{F}(\varphi)\geq\int_{\Omega}\bm{\xi}\cdot(\nabla\phi-\nabla\varphi)\,\mathrm{dx}=\langle v,\phi-\varphi\rangle_{H^{1}}\quad\forall\phi\in H^{1}(\Omega).

This shows v(φ)v\in\partial\mathcal{F}(\varphi) and hence the inclusion 𝒢(φ)(φ)\mathcal{G}(\varphi)\subset\partial\mathcal{F}(\varphi).

For the converse inclusion, we show 𝒢:H1(Ω)2H1(Ω)\mathcal{G}:H^{1}(\Omega)\to 2^{H^{1}(\Omega)^{*}} is a maximal monotone operator, and as \partial\mathcal{F} is maximal monotone we obtain by definition 𝒢(φ)=(φ)\mathcal{G}(\varphi)=\partial\mathcal{F}(\varphi) for all φD()\varphi\in D(\partial\mathcal{F}). Thus, it suffices to study the solvability of the following problem: For fH1(Ω)f\in H^{1}(\Omega)^{*}, find a pair (φ,𝝃)H1(Ω)×L2(Ω,d)(\varphi,\bm{\xi})\in H^{1}(\Omega)\times L^{2}(\Omega,\mathbb{R}^{d}) such that 𝝃A(φ)\bm{\xi}\in\partial A(\nabla\varphi) almost everywhere in Ω\Omega and

φdiv𝝃=f in H1(Ω).\displaystyle\varphi-\mathrm{div}\bm{\xi}=f\quad\text{ in }H^{1}(\Omega)^{*}. (3.15)

This result can be obtained by following a strategy outlined in [12, Thm. 2.17]. First, from recalling the properties of AA in (3.13), we deduce that for 𝝃=(ξ1,,ξd)A(𝒒)\bm{\xi}=(\xi_{1},\dots,\xi_{d})^{\top}\in\partial A(\bm{q}),

𝝃𝒒A(𝒒)c|𝒒|2,|ξi|C(1+|𝒒|),\displaystyle\bm{\xi}\cdot\bm{q}\geq A(\bm{q})\geq c|\bm{q}|^{2},\quad|\xi_{i}|\leq C\big{(}1+|\bm{q}|\big{)}, (3.16)

where the first inequality is obtained from applying the relation A(𝒓)A(𝒒)𝝃(𝒓𝒒)A(\bm{r})-A(\bm{q})\geq\bm{\xi}\cdot(\bm{r}-\bm{q}) with the choice 𝒓=𝟎\bm{r}=\bm{0}, and the second inequality is obtained from the second property of AA in (3.13) with the choice 𝒓=𝒒+𝒆i\bm{r}=\bm{q}+{\bm{e}}_{i} where 𝒆i{\bm{e}}_{i} is the canonical ii-th unit vector in d\mathbb{R}^{d}. Taking the maximum over i{1,,d}i\in\{1,\dots,d\} and then the supremum over all elements 𝝃A(𝒒)\bm{\xi}\in\partial A(\bm{q}), the second inequality in (3.16) implies

sup{|𝝃|:𝝃A(𝒒)}C(1+|𝒒|).\displaystyle\sup\Big{\{}|\bm{\xi}|\,:\,\bm{\xi}\in\partial A(\bm{q})\Big{\}}\leq C\big{(}1+|\bm{q}|\big{)}. (3.17)

Next, for λ>0\lambda>0, we introduce the Moreau–Yosida approximation BλB_{\lambda} of B:=AB:=\partial A as

Bλ(𝒔)=1λ(𝒔(I+λB)1𝒔),𝒔d,\displaystyle B_{\lambda}({\bm{s}})=\frac{1}{\lambda}\Big{(}{\bm{s}}-(I+\lambda B)^{-1}{\bm{s}}\Big{)},\quad{\bm{s}}\in\mathbb{R}^{d},

where II denotes the identity map. The maximal monotonicity of BB guarantees that (I+λB)1(I+\lambda B)^{-1} is a well-defined operator and it is bounded independently of λ\lambda (see [35, p. 524, Thms. 1 and 2]). It is well-known that BλB_{\lambda} is single-valued, Lipschitz continuous with Lipschitz constant 1λ\frac{1}{\lambda}, and Bλ(𝒔)B_{\lambda}({\bm{s}}) is an element of B((I+λB)1𝒔)B((I+\lambda B)^{-1}{\bm{s}}). Then, by (3.16) and (3.17) (cf. [12, p. 84, (2.143)-(2.144)]), we infer from the identity 𝒒=(I+λB)1𝒒+λBλ(𝒒)\bm{q}=(I+\lambda B)^{-1}\bm{q}+\lambda B_{\lambda}(\bm{q}) that for all 𝒒d\bm{q}\in\mathbb{R}^{d} and λ(0,λ0)\lambda\in(0,\lambda_{0}), λ0:=1c\lambda_{0}:=\frac{1}{c} with constant cc from (3.13),

|Bλ(𝒒)|\displaystyle|B_{\lambda}(\bm{q})| C(1+|(I+λB)1𝒒|)C(1+|𝒒|),\displaystyle\leq C\big{(}1+|(I+\lambda B)^{-1}\bm{q}|\big{)}\leq C\big{(}1+|\bm{q}|\big{)}, (3.18a)
Bλ(𝒒)𝒒\displaystyle B_{\lambda}(\bm{q})\cdot\bm{q} =Bλ(𝒒)(I+λB)1𝒒+λ|Bλ(𝒒)|2c|(I+λB)1𝒒|2+λ|Bλ(𝒒)|2\displaystyle=B_{\lambda}(\bm{q})\cdot(I+\lambda B)^{-1}\bm{q}+\lambda|B_{\lambda}(\bm{q})|^{2}\geq c|(I+\lambda B)^{-1}\bm{q}|^{2}+\lambda|B_{\lambda}(\bm{q})|^{2}
c2|𝒒|2+λ(1cλ)|Bλ(𝒒)|2c2|𝒒|2,\displaystyle\geq\frac{c}{2}|\bm{q}|^{2}+\lambda(1-c\lambda)|B_{\lambda}(\bm{q})|^{2}\geq\frac{c}{2}|\bm{q}|^{2}, (3.18b)

with constants independent of λ\lambda. We now consider the approximation problem: For fH1(Ω)f\in H^{1}(\Omega)^{*}, λ(0,λ0),\lambda\in(0,\lambda_{0}), find φλH1(Ω)\varphi_{\lambda}\in H^{1}(\Omega) such that

Tλφλ,ζH1:=Ωφλζ+Bλ(φλ)ζ+λφλζdx=f,ζH1ζH1(Ω).\displaystyle\langle T_{\lambda}\varphi_{\lambda},\zeta\rangle_{H^{1}}:=\int_{\Omega}\varphi_{\lambda}\zeta+B_{\lambda}(\nabla\varphi_{\lambda})\cdot\nabla\zeta+\lambda\nabla\varphi_{\lambda}\cdot\nabla\zeta\,\mathrm{dx}=\langle f,\zeta\rangle_{H^{1}}\quad\forall\zeta\in H^{1}(\Omega). (3.19)

It is not difficult to verify that TλT_{\lambda} is monotone, demicontinuous (i.e., φnφ\varphi_{n}\to\varphi in H1(Ω)H^{1}(\Omega) implies TλφnTλφT_{\lambda}\varphi_{n}\rightharpoonup T_{\lambda}\varphi in H1(Ω)H^{1}(\Omega)^{*} as nn\to\infty), and coercive (i.e., Tλφ,φH1c0φH1\langle T_{\lambda}\varphi,\varphi\rangle_{H^{1}}\geq c_{0}\|\varphi\|_{H^{1}} for some c0>0c_{0}>0 independent of λ\lambda), and so by standard results (see, e.g., [12, p. 37, Cor. 2.3]) there exists at least one solution φλH1(Ω)\varphi_{\lambda}\in H^{1}(\Omega) to Tλφλ=fT_{\lambda}\varphi_{\lambda}=f in H1(Ω)H^{1}(\Omega)^{*}.

We then establish uniform estimates for {φλ}λ(0,λ0)\{\varphi_{\lambda}\}_{\lambda\in(0,\lambda_{0})} whose weak limit in H1(Ω)H^{1}(\Omega) will be a solution to (3.15), thereby verifying the maximal monotonicity of 𝒢\mathcal{G} and hence completing the proof. Choosing ζ=φλ\zeta=\varphi_{\lambda} in (3.19) and recalling the identity 𝒒=(I+λB)1𝒒+λBλ(𝒒)\bm{q}=(I+\lambda B)^{-1}\bm{q}+\lambda B_{\lambda}(\bm{q}), we obtain

φλ2+λ(Bλ(φλ)2+φλ2)\displaystyle\|\varphi_{\lambda}\|^{2}+\lambda\Big{(}\|B_{\lambda}(\nabla\varphi_{\lambda})\|^{2}+\|\nabla\varphi_{\lambda}\|^{2}\Big{)} (3.20)
+ΩBλ(φλ)(I+λB)1φλdx=f,φλH1.\displaystyle\quad+\int_{\Omega}B_{\lambda}(\nabla\varphi_{\lambda})\cdot(I+\lambda B)^{-1}\nabla\varphi_{\lambda}\,\mathrm{dx}=\langle f,\varphi_{\lambda}\rangle_{H^{1}}.

From (3.18) we immediately infer

φλ2+c2φλ2fH1(Ω)φλH1(Ω)CfH1(Ω)2+min(c4,12)φλH1(Ω)2,\displaystyle\|\varphi_{\lambda}\|^{2}+\frac{c}{2}\|\nabla\varphi_{\lambda}\|^{2}\leq\|f\|_{H^{1}(\Omega)^{*}}\|\varphi_{\lambda}\|_{H^{1}(\Omega)}\leq C\|f\|_{H^{1}(\Omega)^{*}}^{2}+\min\big{(}\tfrac{c}{4},\tfrac{1}{2}\big{)}\|\varphi_{\lambda}\|_{H^{1}(\Omega)}^{2},

which provides a uniform estimate for φλ\varphi_{\lambda} in H1(Ω)H^{1}(\Omega) with respect to λ\lambda. On the other hand, also from (3.18), particularly Bλ(𝒒)𝒒c|(I+λB)1𝒒|2B_{\lambda}(\bm{q})\cdot\bm{q}\geq c|(I+\lambda B)^{-1}\bm{q}|^{2}, we observe that

Bλ(φλ)2C(1+(I+λB)1φλ2)\displaystyle\|B_{\lambda}(\nabla\varphi_{\lambda})\|^{2}\leq C\big{(}1+\|(I+\lambda B)^{-1}\nabla\varphi_{\lambda}\|^{2}\big{)} C,\displaystyle\leq C,
(I+λBλ)1φλφλ2=λ2Bλ(φλ)2\displaystyle\|(I+\lambda B_{\lambda})^{-1}\nabla\varphi_{\lambda}-\nabla\varphi_{\lambda}\|^{2}=\lambda^{2}\|B_{\lambda}(\nabla\varphi_{\lambda})\|^{2} Cλ2,\displaystyle\leq C\lambda^{2},

which implies, along a non-relabeled subsequence λ0\lambda\to 0,

φλ\displaystyle\varphi_{\lambda} φ\displaystyle\rightharpoonup\varphi\quad in H1(Ω),\displaystyle\text{ in }H^{1}(\Omega), (3.21a)
(I+λB)1φλ\displaystyle(I+\lambda B)^{-1}\nabla\varphi_{\lambda} φ\displaystyle\rightharpoonup\nabla\varphi\quad in L2(Ω,d),\displaystyle\text{ in }L^{2}(\Omega,\mathbb{R}^{d}), (3.21b)
Bλ(φλ)\displaystyle B_{\lambda}(\nabla\varphi_{\lambda}) 𝝃\displaystyle\rightharpoonup\bm{\xi}\quad in L2(Ω,d),\displaystyle\text{ in }L^{2}(\Omega,\mathbb{R}^{d}), (3.21c)

for some limit functions φ\varphi and 𝝃\bm{\xi} satisfying

Ωφζ+𝝃ζdx=f,ζH1ζH1(Ω).\displaystyle\int_{\Omega}\varphi\zeta+\bm{\xi}\cdot\nabla\zeta\,\mathrm{dx}=\langle f,\zeta\rangle_{H^{1}}\quad\forall\zeta\in H^{1}(\Omega).

To show that 𝝃(𝒙)A(φ(𝒙))\bm{\xi}(\bm{x})\in\partial A(\nabla\varphi(\bm{x})) for almost every 𝒙Ω\bm{x}\in\Omega, for arbitrary ϕD(A){\bm{\phi}}\in D(\partial A) and 𝜻B(ϕ)L2(Ω,d){\bm{\zeta}}\in B({\bm{\phi}})\subset L^{2}(\Omega,\mathbb{R}^{d}), we use the inequality

Ω(Bλ(φλ)𝜻)((I+λB)1φλϕ)dx0\displaystyle\int_{\Omega}(B_{\lambda}(\nabla\varphi_{\lambda})-{\bm{\zeta}})\cdot((I+\lambda B)^{-1}\nabla\varphi_{\lambda}-{\bm{\phi}})\,\mathrm{dx}\geq 0

obtained from the relation Bλ(φλ)B((I+λB)1φλ)B_{\lambda}(\nabla\varphi_{\lambda})\in B((I+\lambda B)^{-1}\nabla\varphi_{\lambda}) and the monotonicity of BB. Then, passing to the limit λ0\lambda\to 0 with (3.21), as well as

ΩBλ(φλ)(I+λB)1φλdx\displaystyle\int_{\Omega}B_{\lambda}(\nabla\varphi_{\lambda})\cdot(I+\lambda B)^{-1}\nabla\varphi_{\lambda}\,\mathrm{dx} =f,φλH1φλ2λφλ2λBλ(φλ)2\displaystyle=\langle f,\varphi_{\lambda}\rangle_{H^{1}}-\|\varphi_{\lambda}\|^{2}-\lambda\|\nabla\varphi_{\lambda}\|^{2}-\lambda\|B_{\lambda}(\nabla\varphi_{\lambda})\|^{2}
f,φH1φ2=Ω𝝃φdx,\displaystyle\to\langle f,\varphi\rangle_{H^{1}}-\|\varphi\|^{2}=\int_{\Omega}\bm{\xi}\cdot\nabla\varphi\,\mathrm{dx},

we arrive at

Ω(𝝃𝜻)(φϕ)dx0.\displaystyle\int_{\Omega}(\bm{\xi}-{\bm{\zeta}})\cdot(\nabla\varphi-{\bm{\phi}})\,\mathrm{dx}\geq 0.

Picking ϕ=(I+B)1(𝝃+φ){\bm{\phi}}=(I+B)^{-1}(\bm{\xi}+\nabla\varphi), so that 𝜻:=𝝃+φϕB(ϕ){\bm{\zeta}}:=\bm{\xi}+\nabla\varphi-{\bm{\phi}}\in B({\bm{\phi}}), we see that the above reduces to

Ω|φϕ|2dx=0,\displaystyle\int_{\Omega}|\nabla\varphi-{\bm{\phi}}|^{2}\,\mathrm{dx}=0,

which implies φ=ϕ\nabla\varphi={\bm{\phi}} and 𝜻=𝝃B(φ)=A(φ){\bm{\zeta}}=\bm{\xi}\in B(\nabla\varphi)=\partial A(\nabla\varphi) almost everywhere in Ω\Omega. This shows that (φ,𝝃)H1(Ω)×L2(Ω,d)(\varphi,\bm{\xi})\in H^{1}(\Omega)\times L^{2}(\Omega,\mathbb{R}^{d}) is a solution to (3.15) concluding the proof. ∎

4 Analysis of the structural optimization problem

By Lemma 2.1 the solution operator 𝑺:L(Ω)HD1(Ω,d)\bm{S}:L^{\infty}(\Omega)\to H^{1}_{D}(\Omega,\mathbb{R}^{d}), 𝑺(φ)=𝒖\bm{S}(\varphi)=\bm{u} where 𝒖\bm{u} is the unique solution to (2.3) corresponding to the design variable φ\varphi, is well-defined. This allows us to consider the reduced functional 𝒥:𝒱m\mathcal{J}:\mathcal{V}_{m}\to\mathbb{R},

𝒥(ϕ):=J(ϕ,𝑺(ϕ))\mathcal{J}(\phi):=J(\phi,\bm{S}(\phi))

in the structural optimization problem (P). Invoking [35, Thm. 1, §8.2.2], the convexity of γ\gamma and hence of A()=12|γ()|2A(\cdot)=\frac{1}{2}|\gamma(\cdot)|^{2} yields that the gradient term in (2.7) is weakly lower semicontinuous in H1(Ω)H^{1}(\Omega). Then, following the proof of [22, Thm. 4.1] we obtain the existence of an optimal design to (P).

Theorem 4.1.

Suppose that (A1)–(A3) hold, and let (𝐟,𝐠)L2(Ω,d)×L2(Γg,d)(\bm{f},\bm{g})\in L^{2}(\Omega,\mathbb{R}^{d})\times L^{2}(\Gamma_{g},\mathbb{R}^{d}). Then, (P) admits a minimiser φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m}.

Proof.

Since the proof is now rather standard, we only sketch some of the essential details. Using the bound (2.5) we infer that the reduced functional 𝒥\mathcal{J} is bounded from below in 𝒱m\mathcal{V}_{m}. This allows us to consider a minimizing sequence {ϕn}n𝒱m\{\phi_{n}\}_{n\in\mathbb{N}}\subset\mathcal{V}_{m} such that 𝒥(ϕn)infζ𝒱m𝒥(ζ)\mathcal{J}(\phi_{n})\to\inf_{\zeta\in\mathcal{V}_{m}}\mathcal{J}(\zeta) as nn\to\infty. Then, again by (2.5) and also (A4) we deduce that {ϕn}n\{\phi_{n}\}_{n\in\mathbb{N}} is uniformly bounded in H1(Ω)L(Ω)H^{1}(\Omega)\cap L^{\infty}(\Omega), from which we extract a non-relabeled subsequence converging weakly to φ¯\overline{\varphi} in H1(Ω)H^{1}(\Omega). As 𝒱m\mathcal{V}_{m} is a convex and closed set, hence weakly sequentially closed, we infer also φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} and by passing to the limit infimum of 𝒥(ϕn)\mathcal{J}(\phi_{n}), employing the weak lower semicontinuity of the gradient term in (2.7) and also the weak convergence 𝑺(ϕn)𝑺(φ¯)\bm{S}(\phi_{n})\to\bm{S}(\overline{\varphi}) in H1(Ω,d)H^{1}(\Omega,\mathbb{R}^{d}) we infer that 𝒥(φ¯)=infζ𝒱m𝒥(ζ)\mathcal{J}(\overline{\varphi})=\inf_{\zeta\in\mathcal{V}_{m}}\mathcal{J}(\zeta), which implies that φ¯\overline{\varphi} is a solution to (P). ∎

To derive the first-order optimality conditions, we first take note of the following result concerning the Fréchet differentiability of the solution operator 𝑺\bm{S}, which can be inferred from [22, Thm. 3.3]:

Lemma 4.1.

Let (𝐟,𝐠)L2(Ω,d)×L2(Γg,d)(\bm{f},\bm{g})\in L^{2}(\Omega,\mathbb{R}^{d})\times L^{2}(\Gamma_{g},\mathbb{R}^{d}) and φ¯L(Ω)\overline{\varphi}\in L^{\infty}(\Omega). Then 𝐒\bm{S} is Fréchet differentiable at φ¯\overline{\varphi} as a mapping from L(Ω)L^{\infty}(\Omega) to HD1(Ω,d)H^{1}_{D}(\Omega,\mathbb{R}^{d}). Moreover, for every ζL(Ω)\zeta\in L^{\infty}(\Omega), 𝐒(φ¯)(L(Ω);HD1(Ω,d))\bm{S}^{\prime}(\overline{\varphi})\in\mathcal{L}(L^{\infty}(\Omega);H^{1}_{D}(\Omega,\mathbb{R}^{d})) and

𝑺(φ¯)[ζ]=𝒘,\displaystyle\bm{S}^{\prime}(\overline{\varphi})[\zeta]={\bm{w}},

where 𝐰HD1(Ω,d){\bm{w}}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}) is the unique solution to the linearized system

Ω𝒞(φ¯)(𝒘):(𝜼)dx\displaystyle\int_{\Omega}\mathcal{C}(\overline{\varphi})\mathcal{E}({\bm{w}}):\mathcal{E}(\bm{\eta})\,\mathrm{dx} =Ω𝒞(φ¯)ζ(𝒖¯):(𝜼)dx+Ω𝕙(φ¯)ζ𝒇𝜼dx\displaystyle=-\int_{\Omega}\mathcal{C}^{\prime}(\overline{\varphi})\zeta\mathcal{E}(\overline{\bm{u}}):\mathcal{E}(\bm{\eta})\,\mathrm{dx}+\int_{\Omega}{\mathbbm{h}}^{\prime}(\overline{\varphi})\zeta\bm{f}\cdot\bm{\eta}\,\mathrm{dx} (4.1)

for all 𝛈HD1(Ω,d),\bm{\eta}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}), with 𝐮¯=𝐒(φ¯)\overline{\bm{u}}=\bm{S}(\overline{\varphi}).

Then, we introduce the adjoint system for the adjoint variable 𝒑\bm{p} associated to φ¯\overline{\varphi}, whose structure is similar to the state equation (2.3) for 𝒖\bm{u}:

div(𝒞(φ¯)(𝒑))\displaystyle-\mathrm{div}(\mathcal{C}(\overline{\varphi})\mathcal{E}(\bm{p})) =β𝕙(φ¯)𝒇\displaystyle=\beta{\mathbbm{h}}(\overline{\varphi})\bm{f}\quad in Ω,\displaystyle\text{ in }\Omega, (4.2a)
𝒑\displaystyle\bm{p} =𝟎\displaystyle=\bm{0}\quad on ΓD,\displaystyle\text{ on }\Gamma_{D}, (4.2b)
(𝒞(φ¯)(𝒑))𝒏\displaystyle(\mathcal{C}(\overline{\varphi})\mathcal{E}(\bm{p}))\bm{n} =β𝒈\displaystyle=\beta\bm{g}\quad on Γg,\displaystyle\text{ on }\Gamma_{g}, (4.2c)
(𝒞(φ¯)(𝒑))𝒏\displaystyle(\mathcal{C}(\overline{\varphi})\mathcal{E}(\bm{p}))\bm{n} =𝟎\displaystyle=\bm{0}\quad on Γ0.\displaystyle\text{ on }\Gamma_{0}. (4.2d)

Via a similar argument (see also [22, Thm. 4.3]), the well-posedness of the adjoint system is straightforward:

Lemma 4.2.

For any φ¯L(Ω)\overline{\varphi}\in L^{\infty}(\Omega) and (𝐟,𝐠)L2(Ω,d)×L2(Γg,d)(\bm{f},\bm{g})\in L^{2}(\Omega,\mathbb{R}^{d})\times L^{2}(\Gamma_{g},\mathbb{R}^{d}), there exists a unique solution 𝐩HD1(Ω,d)\bm{p}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}) to (4.2) satisfying

Ω𝒞(φ¯)(𝒑):(𝒗)dx=βΩ𝕙(φ¯)𝒇𝒗dx+βΓg𝒈𝒗dd1\displaystyle\int_{\Omega}\mathcal{C}(\overline{\varphi})\mathcal{E}(\bm{p}):\mathcal{E}(\bm{v})\,\mathrm{dx}=\beta\int_{\Omega}{\mathbbm{h}}(\overline{\varphi})\bm{f}\cdot\bm{v}\,\mathrm{dx}+\beta\int_{\Gamma_{g}}\bm{g}\cdot\bm{v}\,\mathrm{d}\mathcal{H}^{d-1} (4.3)

for all 𝐯HD1(Ω,d)\bm{v}\in H^{1}_{D}(\Omega,\mathbb{R}^{d}). In fact, if 𝐮¯=𝐒(φ¯)HD1(Ω,d)\overline{\bm{u}}=\bm{S}(\overline{\varphi})\in H^{1}_{D}(\Omega,\mathbb{R}^{d}) is the unique solution to (2.4), then 𝐩=β𝐮¯\bm{p}=\beta\overline{\bm{u}}.

Our main result is the following first-order optimality conditions for the structural optimization problem with anisotropy (P):

Theorem 4.2.

Suppose (A1)–(A3) hold. Let (𝐟,𝐠)L2(Ω,d)×L2(Γg,d)(\bm{f},\bm{g})\in L^{2}(\Omega,\mathbb{R}^{d})\times L^{2}(\Gamma_{g},\mathbb{R}^{d}) and φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} be an optimal design variable with associated state 𝐮¯=𝐒(φ¯)\overline{\bm{u}}=\bm{S}(\overline{\varphi}). Then, there exists 𝛏A(φ¯)\bm{\xi}\in\partial A(\nabla\overline{\varphi}) almost everywhere in Ω\Omega such that

α^Ωε𝝃(ϕφ¯)+1εΨ0(φ¯)(ϕφ¯)dx\displaystyle\widehat{\alpha}\int_{\Omega}\varepsilon\bm{\xi}\cdot\nabla(\phi-\overline{\varphi})+\frac{1}{\varepsilon}\Psi^{\prime}_{0}(\overline{\varphi})(\phi-\overline{\varphi})\,\mathrm{dx} (4.4)
+βΩ2𝕙(φ¯)(ϕφ¯)𝒇𝒖¯𝒞(φ¯)(ϕφ¯)(𝒖¯):(𝒖¯)dx0\displaystyle\quad+\beta\int_{\Omega}2{\mathbbm{h}}^{\prime}(\overline{\varphi})(\phi-\overline{\varphi})\bm{f}\cdot\overline{\bm{u}}-\mathcal{C}^{\prime}(\overline{\varphi})(\phi-\overline{\varphi})\mathcal{E}(\overline{\bm{u}}):\mathcal{E}(\overline{\bm{u}})\,\mathrm{dx}\geq 0

for all ϕ𝒱m\phi\in\mathcal{V}_{m}.

Proof.

Recalling the definition of the functional \mathcal{F} in (3.14), we introduce

𝒦(ϕ):=Ωα^εΨ0(ϕ)+β𝕙(ϕ)𝒇𝑺(ϕ)dx+βΓg𝒈𝑺(ϕ)dd1,\displaystyle\mathcal{K}(\phi):=\int_{\Omega}\frac{\widehat{\alpha}}{\varepsilon}\Psi_{0}(\phi)+\beta{\mathbbm{h}}(\phi)\bm{f}\cdot\bm{S}(\phi)\,\mathrm{dx}+\beta\int_{\Gamma_{g}}\bm{g}\cdot\bm{S}(\phi)\,\mathrm{d}\mathcal{H}^{d-1}, (4.5)

so that the reduced functional 𝒥\mathcal{J} can be expressed as 𝒥(ϕ)=α^ε(ϕ)+𝒦(ϕ)\mathcal{J}(\phi)={\widehat{\alpha}}\varepsilon\mathcal{F}(\phi)+\mathcal{K}(\phi). Using the differentiability of 𝑺\bm{S} we infer that 𝒦\mathcal{K} is also Fréchet differentiable with derivative at φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} in direction ζ𝒱={fH1(Ω):f(𝒙)[1,1] a.e. in Ω}\zeta\in\mathcal{V}=\{f\in H^{1}(\Omega)\,:\,f({\bm{x}})\in[-1,1]\text{ a.e.~{}in }\Omega\} given as

𝒦(φ¯)[ζ]=Ωα^εΨ0(φ¯)ζ+β𝕙(φ¯)ζ𝒇𝒖¯+β𝕙(φ¯)𝒇𝒘dx+βΓg𝒈𝒘dd1,\displaystyle\mathcal{K}^{\prime}(\overline{\varphi})[\zeta]=\int_{\Omega}\frac{\widehat{\alpha}}{\varepsilon}\Psi_{0}^{\prime}(\overline{\varphi})\zeta+\beta{\mathbbm{h}}^{\prime}(\overline{\varphi})\zeta\bm{f}\cdot\overline{\bm{u}}+\beta{\mathbbm{h}}(\overline{\varphi})\bm{f}\cdot{\bm{w}}\,\mathrm{dx}+\beta\int_{\Gamma_{g}}\bm{g}\cdot{\bm{w}}\,\mathrm{d}\mathcal{H}^{d-1}, (4.6)

where 𝒘=𝑺(φ¯)[ζ]{\bm{w}}=\bm{S}^{\prime}(\overline{\varphi})[\zeta] is unique solution to the linearized system (4.1) and 𝒖¯=𝑺(φ¯)\overline{\bm{u}}=\bm{S}(\overline{\varphi}). We can simplify this using the adjoint system by testing (4.3) with 𝒗=𝒘\bm{v}={\bm{w}} and testing (4.1) with 𝜼=𝒑\bm{\eta}=\bm{p} to obtain the identity

βΩ𝕙(φ¯)𝒇𝒘dx+βΓg𝒈𝒘dd1=Ω𝕙(φ¯)ζ𝒇𝒑𝒞(φ¯)ζ(𝒖¯):(𝒑)dx,\beta\int_{\Omega}{\mathbbm{h}}(\overline{\varphi})\bm{f}\cdot{\bm{w}}\,\mathrm{dx}+\beta\int_{\Gamma_{g}}\bm{g}\cdot{\bm{w}}\,\mathrm{d}\mathcal{H}^{d-1}=\int_{\Omega}{\mathbbm{h}}^{\prime}(\overline{\varphi})\zeta\bm{f}\cdot\bm{p}-\mathcal{C}^{\prime}(\overline{\varphi})\zeta\mathcal{E}(\overline{\bm{u}}):\mathcal{E}(\bm{p})\,\mathrm{dx},

so that, together with the relation 𝒑=β𝒖¯\bm{p}=\beta\overline{\bm{u}}, (4.6) becomes

𝒦(φ¯)[ζ]=Ωα^εΨ0(φ¯)ζ+2β𝕙(φ¯)ζ𝒇𝒖¯β𝒞(φ¯)ζ(𝒖¯):(𝒖¯)dx.\displaystyle\mathcal{K}^{\prime}(\overline{\varphi})[\zeta]=\int_{\Omega}\frac{\widehat{\alpha}}{\varepsilon}\Psi_{0}^{\prime}(\overline{\varphi})\zeta+2\beta{\mathbbm{h}}^{\prime}(\overline{\varphi})\zeta\bm{f}\cdot\overline{\bm{u}}-\beta\mathcal{C}^{\prime}(\overline{\varphi})\zeta\mathcal{E}(\overline{\bm{u}}):\mathcal{E}(\overline{\bm{u}})\,\mathrm{dx}. (4.7)

On the other hand, the convexity of \mathcal{F} and the optimality of φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} imply that

0\displaystyle 0 𝒥(φ¯+t(ϕφ¯))𝒥(φ¯)\displaystyle\leq\mathcal{J}(\overline{\varphi}+t(\phi-\overline{\varphi}))-\mathcal{J}(\overline{\varphi})
=α^ε((1t)φ¯+tϕ)α^ε(φ¯)+𝒦(φ¯+t(ϕφ¯))𝒦(φ¯)\displaystyle=\widehat{\alpha}\varepsilon\mathcal{F}((1-t)\overline{\varphi}+t\phi)-\widehat{\alpha}\varepsilon\mathcal{F}(\overline{\varphi})+\mathcal{K}(\overline{\varphi}+t(\phi-\overline{\varphi}))-\mathcal{K}(\overline{\varphi})
tα^ε((ϕ)(φ¯))+𝒦(φ¯+t(ϕφ¯))𝒦(φ¯)\displaystyle\leq t\widehat{\alpha}\varepsilon(\mathcal{F}(\phi)-\mathcal{F}(\overline{\varphi}))+\mathcal{K}(\overline{\varphi}+t(\phi-\overline{\varphi}))-\mathcal{K}(\overline{\varphi})

holds for all t[0,1]t\in[0,1] and arbitrary ϕ𝒱m\phi\in\mathcal{V}_{m}. Dividing by tt and passing to the limit t0t\to 0 yields

0α^ε(ϕ)α^ε(φ¯)+𝒦(φ¯)[ϕφ¯]ϕ𝒱m.\displaystyle 0\leq\widehat{\alpha}\varepsilon\mathcal{F}(\phi)-\widehat{\alpha}\varepsilon\mathcal{F}(\overline{\varphi})+\mathcal{K}^{\prime}(\overline{\varphi})[\phi-\overline{\varphi}]\quad\forall\phi\in\mathcal{V}_{m}. (4.8)

The above inequality allows us to interpret φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} as a solution to the convex minimization problem

minζH1(Ω)(α^ε(ζ)+𝒦(φ¯)[ζ]+𝕀𝒱m(ζ)) where 𝕀𝒱m(ζ)={0 if ζ𝒱m,+ otherwise,\displaystyle\min_{\zeta\in H^{1}(\Omega)}\Big{(}\widehat{\alpha}\varepsilon\mathcal{F}(\zeta)+\mathcal{K}^{\prime}(\overline{\varphi})[\zeta]+\mathbb{I}_{\mathcal{V}_{m}}(\zeta)\Big{)}\quad\text{ where }\mathbb{I}_{\mathcal{V}_{m}}(\zeta)=\begin{cases}0&\text{ if }\zeta\in\mathcal{V}_{m},\\ +\infty&\text{ otherwise},\end{cases} (4.9)

denotes the indicator function of the set 𝒱m\mathcal{V}_{m}. Using the well-known sum rule for subdifferentials of convex functions, see, e.g., [13, Cor. 2.63], the inequality (4.8) can be interpreted as

0(α^ε+𝒦(φ¯)+𝕀𝒱)(φ¯)=α^ε(φ¯)+{𝒦(φ¯)}+𝕀𝒱m(φ¯),0\in\partial\big{(}\widehat{\alpha}\varepsilon\mathcal{F}+\mathcal{K}^{\prime}(\overline{\varphi})+\mathbb{I}_{\mathcal{V}}\big{)}(\overline{\varphi})=\widehat{\alpha}\varepsilon\partial\mathcal{F}(\overline{\varphi})+\{\mathcal{K}^{\prime}(\overline{\varphi})\}+\partial\mathbb{I}_{\mathcal{V}_{m}}(\overline{\varphi}),

where \partial denotes the subdifferential mapping in H1(Ω)H^{1}(\Omega). This implies the existence of elements v(φ¯)v\in\partial\mathcal{F}(\overline{\varphi}) and ψ𝕀𝒱m(φ¯)\psi\in\partial\mathbb{I}_{\mathcal{V}_{m}}(\overline{\varphi}) such that

0=α^εv+𝒦(φ¯)+ψ.\displaystyle 0=\widehat{\alpha}\varepsilon v+\mathcal{K}^{\prime}(\overline{\varphi})+\psi. (4.10)

For arbitrary ϕ𝒱m\phi\in\mathcal{V}_{m}, by definition of 𝕀𝒱m(φ¯)\partial\mathbb{I}_{\mathcal{V}_{m}}(\overline{\varphi}) we have

𝕀𝒱m(ϕ)𝕀𝒱m(φ¯)=0ψ,ϕφ¯H1α^εv,ϕφ¯H1+𝒦(φ¯)[ϕφ¯]0.\mathbb{I}_{\mathcal{V}_{m}}(\phi)-\mathbb{I}_{\mathcal{V}_{m}}(\overline{\varphi})=0\geq\langle\psi,\phi-\overline{\varphi}\rangle_{H^{1}}\quad\implies\quad\langle\widehat{\alpha}\varepsilon v,\phi-\overline{\varphi}\rangle_{H^{1}}+\mathcal{K}^{\prime}(\overline{\varphi})[\phi-\overline{\varphi}]\geq 0.

Then, using Lemma 3.1, there exists 𝝃L2(Ω,d)\bm{\xi}\in L^{2}(\Omega,\mathbb{R}^{d}) with 𝝃A(φ¯)\bm{\xi}\in\partial A(\nabla\overline{\varphi}) almost everywhere in Ω\Omega and

(α^ε𝝃,(ϕφ¯))+𝒦(φ¯)[ϕφ¯]0ϕ𝒱m.(\widehat{\alpha}\varepsilon\bm{\xi},\nabla(\phi-\overline{\varphi}))+\mathcal{K}^{\prime}(\overline{\varphi})[\phi-\overline{\varphi}]\geq 0\quad\forall\phi\in\mathcal{V}_{m}.

The optimality condition (4.4) is then a consequence of the above and (4.7) with ζ=ϕφ¯\zeta=\phi-\overline{\varphi}. ∎

Remark 4.1.

An alternate formulation of the optimality condition (4.4) is as follows: There exist 𝛏A(φ¯)\bm{\xi}\in\partial A(\nabla\overline{\varphi}), θ𝕀[1,1](φ¯)\theta\in\partial\mathbb{I}_{[-1,1]}(\overline{\varphi}), and a Lagrange multiplier μ\mu\in\mathbb{R} for the mass constraint such that

α^Ωε𝝃ζ+1εΨ0(φ¯)ζ+θζdx+μΩζdx\displaystyle\widehat{\alpha}\int_{\Omega}\varepsilon\bm{\xi}\cdot\nabla\zeta+\frac{1}{\varepsilon}\Psi^{\prime}_{0}(\overline{\varphi})\zeta+\theta\zeta\,\mathrm{dx}+\mu\int_{\Omega}\zeta\,\mathrm{dx} (4.11)
+βΩ2𝕙(φ¯)ζ𝒇𝒖¯𝒞(φ¯)ζ(𝒖¯):(𝒖¯)dx=0\displaystyle\quad+\beta\int_{\Omega}2{\mathbbm{h}}^{\prime}(\overline{\varphi})\zeta\bm{f}\cdot\overline{\bm{u}}-\mathcal{C}^{\prime}(\overline{\varphi})\zeta\mathcal{E}(\overline{\bm{u}}):\mathcal{E}(\overline{\bm{u}})\,\mathrm{dx}=0

for all ζH1(Ω)\zeta\in H^{1}(\Omega). In the above, the subdifferential of the indicator function 𝕀[1,1]\mathbb{I}_{[-1,1]} has the explicit characterization

𝕀[1,1](s)={[0,) if s=1,0 if |s|<1,(,0] if s=1.\partial\mathbb{I}_{[-1,1]}(s)=\begin{cases}[0,\infty)&\text{ if }s=1,\\ 0&\text{ if }|s|<1,\\ (-\infty,0]&\text{ if }s=-1.\end{cases}

Indeed, instead of (4.9) we can interpret φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} as a solution to the minimization problem

minζH1(Ω)(α^ε(ζ)+𝒦(φ¯)[ζ]+𝕀[1,1](ζ)+𝕀m(ζ)),𝕀m(ζ)={0 if Ωζdx=m|Ω|,+ otherwise,\min_{\zeta\in H^{1}(\Omega)}\Big{(}\widehat{\alpha}\varepsilon\mathcal{F}(\zeta)+\mathcal{K}^{\prime}(\overline{\varphi})[\zeta]+\mathbb{I}_{[-1,1]}(\zeta)+\mathbb{I}_{m}(\zeta)\Big{)},\quad\mathbb{I}_{m}(\zeta)=\begin{cases}0&\text{ if }\int_{\Omega}\zeta\,\mathrm{dx}=m|\Omega|,\\ +\infty&\text{ otherwise},\end{cases}

which then yields the existence of elements θ𝕀[1,1](φ¯)\theta\in\partial\mathbb{I}_{[-1,1]}(\overline{\varphi}) and η𝕀m(φ¯)\eta\in\partial\mathbb{I}_{m}(\overline{\varphi}) such that

0=α^εv+𝒦(φ¯)+θ+η.0=\widehat{\alpha}\varepsilon v+\mathcal{K}^{\prime}(\overline{\varphi})+\theta+\eta.

For arbitrary ζH1(Ω)\zeta\in H^{1}(\Omega), we test the above equality with ζ(ζ)Ω\zeta-(\zeta)_{\Omega}, where (ζ)Ω=1|Ω|Ωζdx(\zeta)_{\Omega}=\frac{1}{|\Omega|}\int_{\Omega}\zeta\,\mathrm{dx}, and on noting that

0=𝕀m(φ¯±(ζ(ζ)Ω))𝕀m(φ¯)η,±(ζ(ζ)Ω)H1η,ζ(ζ)ΩH1=0,0={\mathbb{I}_{m}}(\overline{\varphi}\pm(\zeta-(\zeta)_{\Omega}))-{\mathbb{I}_{m}}(\overline{\varphi})\geq\langle\eta,\pm(\zeta-(\zeta)_{\Omega})\rangle_{H^{1}}\quad\implies\quad\langle\eta,\zeta-(\zeta)_{\Omega}\rangle_{H^{1}}=0,

this yields (4.11) with Lagrange multiplier μ:=|Ω|1(𝒦(φ¯)[1]+θ)\mu:=-|\Omega|^{-1}(\mathcal{K}^{\prime}(\overline{\varphi})[1]+\theta).

5 Sharp interface asymptotics

Our interest is to study the behavior of solutions under the sharp interface limit ε0\varepsilon\to 0, which connects our phase field approach with the shape optimization approach of [4].

5.1 Γ\Gamma-convergence of the anisotropic Ginzburg–Landau functional

We begin with the Γ\Gamma-convergence of the extended anisotropic Ginzburg–Landau functional (3.4a) to the extended anisotropic perimeter functional (3.4b), which is formulated as follows:

Lemma 5.1.

Let Ωd\Omega\subset\mathbb{R}^{d} be an open bounded domain with Lipschitz boundary. Let γ:d\gamma:\mathbb{R}^{d}\to\mathbb{R} satisfy (A1)–(A3), Ψ:0\Psi:\mathbb{R}\to\mathbb{R}_{\geq 0} is a double well potential with minima at ±1\pm 1 and define cΨ=112Ψ(s)𝑑sc_{\Psi}=\int_{-1}^{1}\sqrt{2\Psi(s)}ds. Then, for any ε>0\varepsilon>0

  1. (i)(i)

    If {uε}ε>0𝒱m\{u_{\varepsilon}\}_{\varepsilon>0}\subset\mathcal{V}_{m} is a sequence such that lim infε0γ,ε(uε)<\liminf_{\varepsilon\to 0}\mathcal{E}_{\gamma,\varepsilon}(u_{\varepsilon})<\infty and uεu0u_{\varepsilon}\to u_{0} strongly in L1(Ω)L^{1}(\Omega), then u0BVm(Ω,{1,1})u_{0}\in\mathrm{BV}_{m}(\Omega,\{-1,1\}) with γ,0(u0)lim infε0γ,ε(uε)\mathcal{E}_{\gamma,0}(u_{0})\leq\liminf_{\varepsilon\to 0}\mathcal{E}_{\gamma,\varepsilon}(u_{\varepsilon}).

  2. (ii)(ii)

    Let u0BV(Ω,{1,1})u_{0}\in\mathrm{BV}(\Omega,\{-1,1\}). Then, there exists a sequence {uε}ε>0\{u_{\varepsilon}\}_{\varepsilon>0} of Lipschitz continuous functions on Ω\Omega such that uε(𝒙)[1,1]u_{\varepsilon}(\bm{x})\in[-1,1] a.e. in Ω\Omega, uεu0u_{\varepsilon}\to u_{0} strongly in L1(Ω)L^{1}(\Omega), Ωuε(𝒙)dx=Ωu0(𝒙)dx\int_{\Omega}u_{\varepsilon}(\bm{x})\,\mathrm{dx}=\int_{\Omega}u_{0}(\bm{x})\,\mathrm{dx} for all ε>0\varepsilon>0, and lim supε0γ,ε(uε)γ,0(u0)\limsup_{\varepsilon\to 0}\mathcal{E}_{\gamma,\varepsilon}(u_{\varepsilon})\leq\mathcal{E}_{\gamma,0}(u_{0}).

  3. (iii)(iii)

    Let {uε}ε>0𝒱m\{u_{\varepsilon}\}_{\varepsilon>0}\subset\mathcal{V}_{m} be a sequence satisfying supε>0γ,ε(uε)<\sup_{\varepsilon>0}\mathcal{E}_{\gamma,\varepsilon}(u_{\varepsilon})<\infty. Then, there exists a non-relabeled subsequence ε0\varepsilon\to 0 and a limit function uu such that uεuu_{\varepsilon}\to u strongly in L1(Ω)L^{1}(\Omega) with γ,0(u)<\mathcal{E}_{\gamma,0}(u)<\infty.

The first and second assertions are known as the liminf and limsup inequalities, respectively, while the third assertion is the compactness property. In the following, we outline how to adapt the proofs of [20, Thms. 3.1 and 3.4] for the Γ\Gamma-convergence result in the multi-phase case. Our present setting corresponds to the case N=2N=2 in their notation.

Proof.

For arbitrary u𝒱mu\in\mathcal{V}_{m}, we introduce the associated vector 𝒘=(w1,w2)\bm{w}=(w_{1},w_{2}) where w1:=12(1u)w_{1}:=\frac{1}{2}(1-u) and w2:=12(1+u)w_{2}:=\frac{1}{2}(1+u). Then, 𝒘\bm{w} take values in the Gibbs simplex Σ={(w1,w2)2:w1,w20,w1+w2=1}\Sigma=\{(w_{1},w_{2})\in\mathbb{R}^{2}\,:\,w_{1},w_{2}\geq 0,\,w_{1}+w_{2}=1\}. Setting 𝜶1:=(0,1)\bm{\alpha}^{1}:=(0,1)^{\top} and 𝜶2:=(1,0)\bm{\alpha}^{2}:=(1,0)^{\top}, we consider a multiple-well potential W:Σ[0,)W:\Sigma\to[0,\infty) satisfying W1(0)={𝜶1,𝜶2}W^{-1}(0)=\{\bm{\alpha}^{1},\bm{\alpha}^{2}\} and

W(12(1u),12(1+u))=Ψ0(u),\displaystyle W{\Big{(}}\frac{1}{2}(1-u),\frac{1}{2}(1+u){\Big{)}}=\Psi_{0}(u), (5.1)

where the latter relation connects WW to Ψ0\Psi_{0} defined in (2.8). Denoting by M2×dM^{2\times d} the set of 2-by-dd matrices, we consider a function f:Σ×M2×d[0,)f:\Sigma\times M^{2\times d}\to[0,\infty) defined as f(𝒘,𝑿)=2|γ(w1𝑿2w2𝑿1)|2f(\bm{w},\bm{X})=2|\gamma(w_{1}\bm{X}_{2}-w_{2}\bm{X}_{1})|^{2}, where γ\gamma satisfies (A1)–(A3), and 𝑿i\bm{X}_{i} denotes the ii-th row of 𝑿\bm{X}. Taking 𝑿=𝒘\bm{X}=\nabla\bm{w} for 𝒘Σ\bm{w}\in\Sigma yields that 𝑿1+𝑿2=w1+w2=0\bm{X}_{1}+\bm{X}_{2}=\nabla w_{1}+\nabla w_{2}=0 and

f(𝒘,𝒘)=2|γ((1w2)w2+w2w2)|2=12|γ(u)|2\displaystyle f(\bm{w},\nabla\bm{w})=2|\gamma((1-w_{2})\nabla w_{2}+w_{2}\nabla w_{2})|^{2}=\frac{1}{2}|\gamma(\nabla u)|^{2} (5.2)

by the relation w2=12u\nabla w_{2}=\frac{1}{2}\nabla u and the one-homogeneity of γ\gamma. Assumptions (A1)–(A3) on γ\gamma ensures the function ff defined above fulfills the corresponding assumptions in [20, p. 80], and thus by [20, Thm. 3.1], the extended functional

Gε(𝒘)={Ωεf(𝒘,𝒘)+1εW(𝒘)dx if 𝒘H1(Ω;Σ),+ elsewhere in L1(Ω;Σ)\displaystyle G_{\varepsilon}(\bm{w})=\begin{cases}\displaystyle\int_{\Omega}\varepsilon f(\bm{w},\nabla\bm{w})+\frac{1}{\varepsilon}W(\bm{w})\,\mathrm{dx}&\text{ if }\bm{w}\in H^{1}(\Omega;\Sigma),\\[10.0pt] +\infty&\text{ elsewhere in }L^{1}(\Omega;\Sigma)\end{cases}

Γ\Gamma-converges in L1(Ω;Σ)L^{1}(\Omega;\Sigma) to a limit functional

𝒢(𝒘)={S𝒘φ(𝜶1,𝜶2,𝝂)dd1 if 𝒘BV(Ω;{𝜶1,𝜶2}),+ otherwise,\displaystyle\mathcal{G}(\bm{w})=\begin{cases}\displaystyle\int_{S_{\bm{w}}}\varphi(\bm{\alpha}^{1},\bm{\alpha}^{2},\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1}&\text{ if }\bm{w}\in\mathrm{BV}(\Omega;\{\bm{\alpha}^{1},\bm{\alpha}^{2}\}),\\[10.0pt] +\infty&\text{ otherwise},\end{cases}

where the definitions of the set S𝒘S_{\bm{w}}, normal 𝝂\bm{\nu} and function φ\varphi can be found in [20, p. 78, p. 81]. It is clear that γ,ε(u)=Gε((12(1u),12(1+u)))\mathcal{E}_{\gamma,\varepsilon}(u)=G_{\varepsilon}((\frac{1}{2}(1-u),\frac{1}{2}(1+u))) for u𝒱mu\in\mathcal{V}_{m}, and our aim is to show γ,0(u)=𝒢((12(1u),12(1+u)))\mathcal{E}_{\gamma,0}(u)=\mathcal{G}((\frac{1}{2}(1-u),\frac{1}{2}(1+u))) for uBV(Ω,{1,1})u\in\mathrm{BV}(\Omega,\{-1,1\}). For a fixed vector 𝝂𝕊d1\bm{\nu}\in\mathbb{S}^{d-1}, let 𝒞λ\mathcal{C}_{\lambda} denote the (d1)(d-1)-dimensional cube of side λ\lambda lying in the orthogonal complement 𝝂\bm{\nu}^{\perp} centered at the origin. Following [20], a function 𝒘:dΣ\bm{w}:\mathbb{R}^{d}\to\Sigma is 𝒞λ\mathcal{C}_{\lambda}-periodic if 𝒘(𝒙+λ𝒆i)=𝒘(𝒙)\bm{w}(\bm{x}+\lambda\bm{e}_{i})=\bm{w}(\bm{x}) for every 𝒙\bm{x} and every i=1,,d1i=1,\dots,d-1, with 𝒆1,,𝒆d1\bm{e}_{1},\dots,\bm{e}_{d-1} as directions of the sides of 𝒞λ\mathcal{C}_{\lambda}. Setting Qλ,𝝂={𝒛+t𝝂:𝒛𝒞λ,t}Q_{\lambda,\infty}^{\bm{\nu}}=\{\bm{z}+t\bm{\nu}\,:\,\bm{z}\in\mathcal{C}_{\lambda},\,t\in\mathbb{R}\} we denote by 𝒳(Qλ,𝝂)\mathcal{X}(Q_{\lambda,\infty}^{\bm{\nu}}) the class of all functions 𝒘Wloc1,2(d;Σ)\bm{w}\in W^{1,2}_{\mathrm{loc}}(\mathbb{R}^{d};\Sigma) which are 𝒞λ\mathcal{C}_{\lambda}-periodic and satisfy lim𝒙,𝝂𝒘(𝒙)=𝜶1\lim_{\langle\bm{x},\bm{\nu}\rangle\to-\infty}\bm{w}(\bm{x})=\bm{\alpha}^{1} and lim𝒙,𝝂+𝒘(𝒙)=𝜶2\lim_{\langle\bm{x},\bm{\nu}\rangle\to+\infty}\bm{w}(\bm{x})=\bm{\alpha}^{2}. Then, by [20, (6)], the integrand in the Γ\Gamma-limit functional 𝒢\mathcal{G} has the representation formula

φ(𝜶1,𝜶2,𝝂)=limλ+inf{λ1dG1(𝒘,Qλ,𝝂):𝒘𝒳(Qλ,𝝂)},\varphi(\bm{\alpha}^{1},\bm{\alpha}^{2},\bm{\nu})=\lim_{\lambda\to+\infty}\inf\Big{\{}\lambda^{1-d}G_{1}(\bm{w},Q_{\lambda,\infty}^{\bm{\nu}})\,:\,\bm{w}\in\mathcal{X}(Q_{\lambda,\infty}^{\bm{\nu}})\Big{\}},

where G1(𝒘,A)=Af(𝒘,𝒘)+W(𝒘)dxG_{1}(\bm{w},A)=\int_{A}f(\bm{w},\nabla\bm{w})+W(\bm{w})\,\mathrm{dx}. To simplify the above expression, for a fixed 𝝂𝕊d1\bm{\nu}\in\mathbb{S}^{d-1}, consider a function u𝒱mu\in\mathcal{V}_{m} of the form u(𝒙)=ψ(𝒙𝝂/γ(𝝂))u(\bm{x})=\psi(\bm{x}\cdot\bm{\nu}/\gamma(\bm{\nu})) for a monotone function ψ:[1,1]{\psi}:\mathbb{R}\to[-1,1] such that lims±ψ(s)=±1\lim_{s\to\pm\infty}\psi(s)=\pm 1 and satisfies ψ′′(s)=Ψ0(ψ(s))\psi^{\prime\prime}(s)=\Psi_{0}^{\prime}(\psi(s)). For the choice Ψ0(s)=12(1s2)\Psi_{0}(s)=\frac{1}{2}(1-s^{2}) we have the explicit solution (known also as the optimal profile)

ψ(s)={1 if sπ2,sin(s) if |s|π2,1 if sπ2.\displaystyle\psi(s)=\begin{cases}1&\text{ if }s\geq\frac{\pi}{2},\\ \sin(s)&\text{ if }|s|\leq\frac{\pi}{2},\\ -1&\text{ if }s\leq-\frac{\pi}{2}.\end{cases} (5.3)

Furthermore, multiplying the equality ψ′′(s)=Ψ0(ψ(s))\psi^{\prime\prime}(s)=\Psi^{\prime}_{0}(\psi(s)) by ψ(s)\psi^{\prime}(s) and integrating yields the so-called equipartition of energy 12|ψ(s)|2=Ψ0(ψ(s))\frac{1}{2}|\psi^{\prime}(s)|^{2}=\Psi_{0}(\psi(s)) for all ss\in\mathbb{R}. Then, it is clear that 𝒘=(12(1u),12(1+u))𝒳(Qλ,𝝂)\bm{w}=(\frac{1}{2}(1-u),\frac{1}{2}(1+u))\in\mathcal{X}(Q_{\lambda,\infty}^{\bm{\nu}}). By (A1), (5.1) and (5.2), as well as Fubini’s theorem, for u(𝒙)=ψ(𝒙𝝂/γ(𝝂))u(\bm{x})={\psi}(\bm{x}\cdot\bm{\nu}/\gamma(\bm{\nu})) we have u(𝒙)=ψ(𝒙𝝂/γ(𝝂))𝝂/γ(𝝂)\nabla u(\bm{x})={\psi}^{\prime}(\bm{x}\cdot\bm{\nu}/\gamma(\bm{\nu}))\bm{\nu}/\gamma(\bm{\nu}), and

G1(𝒘,Qλ,𝝂)\displaystyle G_{1}(\bm{w},Q_{\lambda,\infty}^{\bm{\nu}}) =𝒞λ12|γ(u(𝒛+t𝝂))|2+Ψ0(u(𝒛+t𝝂))dtd𝒛\displaystyle=\int_{\mathcal{C}_{\lambda}}\int_{\mathbb{R}}\frac{1}{2}|\gamma(\nabla u(\bm{z}+t\bm{\nu}))|^{2}+\Psi_{0}(u(\bm{z}+t\bm{\nu}))\,\mathrm{d}t\,\mathrm{d}\bm{z}
=𝒞λ12|ψ(t/γ(𝝂))|2+Ψ0(ψ(t/γ(𝝂)))dtd𝒛\displaystyle=\int_{\mathcal{C}_{\lambda}}\int_{\mathbb{R}}\frac{1}{2}|\psi^{\prime}(t/\gamma(\bm{\nu}))|^{2}+\Psi_{0}(\psi(t/\gamma(\bm{\nu})))\,\mathrm{d}t\,\mathrm{d}\bm{z}
=|𝒞λ|γ(𝝂)1112|ψ(s)|2+Ψ0(ψ(s))ds=λd1γ(𝝂)112Ψ0(r))dr\displaystyle=|\mathcal{C}_{\lambda}|\gamma(\bm{\nu})\int_{-1}^{1}\frac{1}{2}|\psi^{\prime}(s)|^{2}+\Psi_{0}(\psi(s))\,\mathrm{d}s=\lambda^{d-1}\gamma(\bm{\nu})\int_{-1}^{1}\sqrt{2\Psi_{0}(r))}\,\mathrm{d}r

after a change of variables r=ψ(s)r=\psi(s) and using the equipartition of energy 12|ψ(s)|2=Ψ0(ψ(s))\frac{1}{2}|\psi^{\prime}(s)|^{2}=\Psi_{0}(\psi(s)). Hence, we obtain the identification φ(𝜶1,𝜶2,𝝂)=cΨγ(𝝂)\varphi(\bm{\alpha}^{1},\bm{\alpha}^{2},\bm{\nu})=c_{\Psi}\gamma(\bm{\nu}) with constant cΨ=112Ψ0(r)dr=πc_{\Psi}=\int_{-1}^{1}\sqrt{2\Psi_{0}(r)}\,\mathrm{d}r=\pi. Then, for 𝒘=(w1,w2)BV(Ω,{𝜶1,𝜶2})\bm{w}=(w_{1},w_{2})\in\mathrm{BV}(\Omega,\{\bm{\alpha}^{1},\bm{\alpha}^{2}\}), we define uBV(Ω,{1,1})u\in\mathrm{BV}(\Omega,\{-1,1\}) via the relation u=w2w1u=w_{2}-w_{1}, and this allows us to identify S𝒘={u=1}S_{\bm{w}}=\partial^{*}\{u=1\} and consequently 𝒢(𝒘)=γ,0(u)\mathcal{G}(\bm{w})=\mathcal{E}_{\gamma,0}(u). Then, the liminf, limsup and compactness properties follow directly from [20, Thms. 3.1 and 3.4]. ∎

5.2 Formally matched asymptotic analysis

In this section we consider an anisotropy function γC2(d)W1,(d)\gamma\in C^{2}(\mathbb{R}^{d})\cap W^{1,\infty}(\mathbb{R}^{d}) satisfying (A1)–(A3) in the structural optimization problem (P), as well as a more regular body force 𝒇H1(Ω,d)\bm{f}\in H^{1}(\Omega,\mathbb{R}^{d}). The differentiability of γ\gamma implies the characterization of the subdifferential A(φ)\partial A(\nabla\varphi) as the singleton set {[γDγ](φ)}\{[\gamma\mathrm{D}\gamma](\nabla\varphi)\}, where

Dγ(𝒒)=(q1γ,q2γ,,qdγ)(𝒒)d,𝒒=(q1,,qd)d.\mathrm{D}\gamma(\bm{q})=(\partial_{q_{1}}\gamma,\partial_{q_{2}}\gamma,\dots,\partial_{q_{d}}\gamma)(\bm{q})\in\mathbb{R}^{d},\quad\bm{q}=(q_{1},\dots,q_{d})\in\mathbb{R}^{d}.

Then, the corresponding optimality condition for a minimizer (φ¯ε,𝒖¯ε)(\overline{\varphi}_{\varepsilon},\overline{\bm{u}}_{\varepsilon}) to (P) becomes

α^Ωε[γDγ](φ¯ε)(ϕφ¯ε)+1εΨ0(φ¯ε)(ϕφ¯ε)dx\displaystyle\widehat{\alpha}\int_{\Omega}\varepsilon[\gamma\mathrm{D}\gamma](\nabla\overline{\varphi}_{\varepsilon})\cdot\nabla(\phi-\overline{\varphi}_{\varepsilon})+\frac{1}{\varepsilon}\Psi_{0}^{\prime}(\overline{\varphi}_{\varepsilon})(\phi-\overline{\varphi}_{\varepsilon})\,\mathrm{dx} (5.4)
+βΩ[2𝕙(φ¯ε)𝒇𝒖¯ε𝒞(φ¯ε)(𝒖¯ε):(𝒖¯ε)](ϕφ¯ε)dx0ϕ𝒱m,\displaystyle\quad+\beta\int_{\Omega}[2{\mathbbm{h}}^{\prime}(\overline{\varphi}_{\varepsilon})\bm{f}\cdot\overline{\bm{u}}_{\varepsilon}-\mathcal{C}^{\prime}(\overline{\varphi}_{\varepsilon})\mathcal{E}(\overline{\bm{u}}_{\varepsilon}):\mathcal{E}(\overline{\bm{u}}_{\varepsilon})](\phi-\overline{\varphi}_{\varepsilon})\,\mathrm{dx}\geq 0\quad\forall\phi\in{\mathcal{V}_{m}},

and its strong formulation reads as (see (4.11))

β(2𝕙(φ¯ε)𝒇𝒖¯ε𝒞(φ¯ε)(𝒖¯ε):(𝒖¯ε))+με\displaystyle\beta\Big{(}2{\mathbbm{h}}^{\prime}(\overline{\varphi}_{\varepsilon})\bm{f}\cdot\overline{\bm{u}}_{\varepsilon}-\mathcal{C}^{\prime}(\overline{\varphi}_{\varepsilon})\mathcal{E}(\overline{\bm{u}}_{\varepsilon}):\mathcal{E}(\overline{\bm{u}}_{\varepsilon})\Big{)}+\mu_{\varepsilon}
+α^(1εΨ0(φ¯ε)εdiv([γDγ](φ¯ε)))\displaystyle\quad+\,\widehat{\alpha}\Big{(}\frac{1}{\varepsilon}\Psi_{0}^{\prime}(\overline{\varphi}_{\varepsilon})-\varepsilon\mathrm{div}\big{(}[\gamma\mathrm{D}\gamma](\nabla\overline{\varphi}_{\varepsilon})\big{)}\Big{)} =0 in Ω{|φ¯ε|<1},\displaystyle=0\quad\text{ in }\Omega\cap\{|\overline{\varphi}_{\varepsilon}|<1\}, (5.5a)
[γDγ](φ¯ε)𝒏\displaystyle[\gamma\mathrm{D}\gamma](\nabla\overline{\varphi}_{\varepsilon})\cdot\bm{n} =0 on Ω,\displaystyle=0\quad\text{ on }\partial\Omega, (5.5b)

with Lagrange multiplier με\mu_{\varepsilon} for the mass constraint. Our aim is to perform a formally matched asymptotic analysis as ε0\varepsilon\to 0, similar as in [25, Sec. 5], in order to infer the sharp interface limit of (5.5). The method proceeds as follows, see [31]: formally we assume that the domain Ω\Omega admits a decomposition Ω=Ωε+ΩεΩεI\Omega=\Omega_{\varepsilon}^{+}\cup\Omega_{\varepsilon}^{-}\cup\Omega_{\varepsilon}^{\rm{I}}, where ΩεI\Omega_{\varepsilon}^{\rm{I}} is an annular domain and φ¯ε𝒱m\overline{\varphi}_{\varepsilon}\in\mathcal{V}_{m} satisfies

φ¯ε=±1 in Ωε±,|φ¯ε|<1 in ΩεI.\displaystyle\overline{\varphi}_{\varepsilon}=\pm 1\text{ in }\Omega_{\varepsilon}^{\pm},\quad|\overline{\varphi}_{\varepsilon}|<1\text{ in }\Omega_{\varepsilon}^{\rm{I}}. (5.6)

The zero level set Λε:={𝒙Ω:φ¯ε(𝒙)=0}\Lambda_{\varepsilon}:=\{\bm{x}\in\Omega\,:\,\overline{\varphi}_{\varepsilon}(\bm{x})=0\} is assumed to converge to a smooth hypersurface Λ\Lambda as ε0\varepsilon\to 0, and that (φ¯ε,𝒖¯ε)(\overline{\varphi}_{\varepsilon},\overline{\bm{u}}_{\varepsilon}) admit an outer expansion in regions in Ωε±\Omega_{\varepsilon}^{\pm} as well as an inner expansion in regions in ΩεI\Omega_{\varepsilon}^{\rm{I}}. We substitute these expansions (in powers of ε\varepsilon) in (2.3) and (5.5), collecting terms of the same order of ε\varepsilon, and with the help of suitable matching conditions connecting these two expansions, we deduce an equation posed on Λ\Lambda. For an introduction and more detailed discussion of this methodology we refer to [37, 47].

To start, let us collect some useful relations. As γ\gamma is positively homogeneous of degree one, taking the relation γ(t𝒑)=tγ(𝒑)\gamma(t\bm{p})=t\gamma(\bm{p}) for t>0t>0 and 𝒑d{𝟎}\bm{p}\in\mathbb{R}^{d}\setminus\{\bm{0}\} and differentiating with respect to tt, and also with respect to 𝒑\bm{p} leads to

Dγ(t𝒑)𝒑=γ(𝒑),Dγ(t𝒑)=Dγ(𝒑),\displaystyle\mathrm{D}\gamma(t\bm{p})\cdot\bm{p}=\gamma(\bm{p}),\quad\mathrm{D}\gamma(t\bm{p})=\mathrm{D}\gamma(\bm{p}), (5.7)

with the latter relation showing that Dγ\mathrm{D}\gamma is positively homogeneous of degree zero. Then, it is easy to see γDγ\gamma\mathrm{D}\gamma is positively homogeneous of degree one. Next, from (3.12), we see that AA is positively homogeneous of degree two. Taking the relation A(t𝒑)=t2A(𝒑)A(t\bm{p})=t^{2}A(\bm{p}) and differentiating with respect to tt twice leads to

D2A(t𝒑)𝒑𝒑=2A(𝒑)[D(γDγ)(𝒑)]𝒑𝒑=|γ(𝒑)|2,\displaystyle\mathrm{D}^{2}A(t\bm{p})\bm{p}\cdot\bm{p}=2A(\bm{p})\quad\implies\quad[\mathrm{D}(\gamma\mathrm{D}\gamma)(\bm{p})]\bm{p}\cdot\bm{p}=|\gamma(\bm{p})|^{2}, (5.8)

where D(γDγ)=γD2γ+DγDγ\mathrm{D}(\gamma\mathrm{D}\gamma)=\gamma\mathrm{D}^{2}\gamma+\mathrm{D}\gamma\otimes\mathrm{D}\gamma with D2γ\mathrm{D}^{2}\gamma denoting the Hessian matrix of γ\gamma. Lastly, differentiating the relation [γDγ](t𝒑)=t[γDγ](𝒑)[\gamma\mathrm{D}\gamma](t\bm{p})=t[\gamma\mathrm{D}\gamma](\bm{p}) with respect to tt and setting t=1t=1 yields

[D(γDγ)(𝒑)]𝒑=(γDγ)(𝒑).\displaystyle[\mathrm{D}(\gamma\mathrm{D}\gamma)(\bm{p})]\bm{p}=(\gamma\mathrm{D}\gamma)(\bm{p}). (5.9)

5.2.1 Outer expansions

For points 𝒙\bm{x} in Ωε±\Omega_{\varepsilon}^{\pm}, we assume an outer expansion of the form

𝒖¯ε(𝒙)=k=0εk𝒖k(𝒙),φ¯ε(𝒙)=k=0εkφk(𝒙),\overline{\bm{u}}_{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\bm{u}_{k}(\bm{x}),\quad\overline{\varphi}_{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\varphi_{k}(\bm{x}),

where all functions are sufficiently smooth and the summations converge. From (5.6) we deduce that

φ¯0(𝒙){1,1},φ¯i(𝒙)=0 for i1.\overline{\varphi}_{0}(\bm{x})\in\{-1,1\},\quad\overline{\varphi}_{i}(\bm{x})=0\text{ for }i\geq 1.

Setting Ω+={φ¯0=1}\Omega_{+}=\{\overline{\varphi}_{0}=1\} and Ω={φ¯0=1}\Omega_{-}=\{\overline{\varphi}_{0}=-1\}, it holds that (Ω+Ω)Ω=Λ(\partial\Omega_{+}\cap\partial\Omega_{-})\cap\Omega=\Lambda. Then, substituting the outer expansion into (2.3), we obtain to order 𝒪(1)\mathcal{O}(1) the following system of equations

div(𝒞(±1)(𝒖0))\displaystyle-\mathrm{div}(\mathcal{C}(\pm 1)\mathcal{E}(\bm{u}_{0})) =𝕙(±1)𝒇\displaystyle={\mathbbm{h}}(\pm 1)\bm{f}  in Ω±,\displaystyle\quad\text{ in }\Omega_{\pm}, (5.10a)
𝒖0\displaystyle\bm{u}_{0} =𝟎\displaystyle=\bm{0}  on ΓDΩ±,\displaystyle\quad\text{ on }\Gamma_{D}\cap\partial\Omega_{\pm}, (5.10b)
(𝒞(±1)(𝒖0))𝒏\displaystyle(\mathcal{C}(\pm 1)\mathcal{E}(\bm{u}_{0}))\bm{n} =𝒈\displaystyle=\bm{g}  on ΓgΩ±,\displaystyle\quad\text{ on }\Gamma_{g}\cap\partial\Omega_{\pm}, (5.10c)
(𝒞(±1)(𝒖0))𝒏\displaystyle(\mathcal{C}(\pm 1)\mathcal{E}(\bm{u}_{0}))\bm{n} =𝟎\displaystyle=\bm{0}  on Γ0Ω±.\displaystyle\quad\text{ on }\Gamma_{0}\cap\partial\Omega_{\pm}. (5.10d)

Moreover, from the definition of the Lagrange multiplier με\mu_{\varepsilon} we see that

με=1εμ1+μ0+𝒪(ε), with μ1:=1|Ω|Ωα^Ψ0(φ¯0)dx=0,\displaystyle\mu_{\varepsilon}=\frac{1}{\varepsilon}\mu_{-1}+\mu_{0}+\mathcal{O}(\varepsilon),\quad\text{ with }\quad\mu_{-1}:=\frac{1}{|\Omega|}\int_{\Omega}\widehat{\alpha}\Psi_{0}^{\prime}(\overline{\varphi}_{0})\,\mathrm{dx}=0, (5.11)
μ0=1|Ω|Ω2β𝕙(φ¯0)𝒇𝒖¯0β𝒞(φ¯0)(𝒖¯0):(𝒖¯0)dx.\displaystyle\mu_{0}=\frac{1}{|\Omega|}\int_{\Omega}2\beta{\mathbbm{h}}^{\prime}(\overline{\varphi}_{0})\bm{f}\cdot\overline{\bm{u}}_{0}-\beta\mathcal{C}^{\prime}(\overline{\varphi}_{0})\mathcal{E}(\overline{\bm{u}}_{0}):\mathcal{E}(\overline{\bm{u}}_{0})\,\mathrm{dx}{.}

It remains to derive the boundary conditions for (5.10) holding on Λ\Lambda, which can be achieved with the inner expansions.

5.2.2 Inner expansions

We assume the outer boundary Γε+\Gamma_{\varepsilon}^{+} and inner boundary Γε\Gamma_{\varepsilon}^{-} of the annular region ΩεI\Omega_{\varepsilon}^{\rm{I}} can be parameterized over the smooth hypersurface Λ\Lambda. Let 𝝂\bm{\nu} denote the unit normal of Λ\Lambda pointing from Ω\Omega_{-} to Ω+\Omega_{+}, and we choose a spatial parameter domain Ud1U\subset\mathbb{R}^{d-1} with a local parameterization 𝒓:Ud\bm{r}:U\to\mathbb{R}^{d} of Λ\Lambda. Let dd denote the signed distance function of Λ\Lambda with d(𝒙)>0d(\bm{x})>0 if 𝒙Ω+\bm{x}\in\Omega_{+}, and denote by z=dεz=\frac{d}{\varepsilon} the rescaled signed distance. Then, by the smoothness of Λ\Lambda, there exists δ0>0\delta_{0}>0 such that for all 𝒙{|d(𝒙)|<δ0}\bm{x}\in\{|d(\bm{x})|<\delta_{0}\}, we have the representation

𝒙=Gε(𝒔,z)=𝒓(𝒔)+εz𝝂(𝒔),\bm{x}=G^{\varepsilon}(\bm{s},z)=\bm{r}(\bm{s})+\varepsilon z\bm{\nu}(\bm{s}),

where 𝒔=(s1,,sd1)U\bm{s}=(s_{1},\dots,s_{d-1})\in U. In particular, 𝒓(𝒔)Λ\bm{r}(\bm{s})\in\Lambda is the projection of 𝒙\bm{x} to Λ\Lambda along the normal direction. This representation allows us to infer the following expansion for gradients and divergences [47]:

𝒙b=1εzb^𝝂+Λb^+𝒪(ε),div𝒙𝒋=1εz𝒋^𝝂+divΛ𝒋^+𝒪(ε)\displaystyle\nabla_{\bm{x}}b=\frac{1}{\varepsilon}\partial_{z}\hat{b}\bm{\nu}+\nabla_{\Lambda}\hat{b}+\mathcal{O}(\varepsilon),\quad\mathrm{div}_{\bm{x}}\bm{j}=\frac{1}{\varepsilon}\partial_{z}\widehat{\bm{j}}\cdot\bm{\nu}+\mathrm{div}_{\Lambda}\widehat{\bm{j}}+\mathcal{O}(\varepsilon) (5.12)

for scalar functions b(𝒙)=b^(𝒔(𝒙),z(𝒙))b(\bm{x})=\hat{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})). In the above Λ\nabla_{\Lambda} is the surface gradient on Λ\Lambda and divΛ\mathrm{div}_{\Lambda} is the surface divergence. Analogously, for a vector function 𝒋\bm{j}, we find that

𝒙𝒋=1εz𝒋^𝝂+Λ𝒋^+𝒪(ε).\nabla_{\bm{x}}\bm{j}=\frac{1}{\varepsilon}\partial_{z}\widehat{\bm{j}}\otimes\bm{\nu}+\nabla_{\Lambda}\widehat{\bm{j}}+\mathcal{O}(\varepsilon).

For points close by Λ\Lambda in ΩεI\Omega_{\varepsilon}^{\rm{I}}, we assume an expansion of the form

𝒖¯ε(𝒙)=k=0εk𝑼k(𝒔,z),φ¯ε(𝒙)=k=0εkΦk(𝒔,z).\overline{\bm{u}}_{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\bm{U}_{k}(\bm{s},z),\quad\overline{\varphi}_{\varepsilon}(\bm{x})=\sum_{k=0}^{\infty}\varepsilon^{k}\Phi_{k}(\bm{s},z).

5.2.3 Matching conditions

In a tubular neighborhood of Λ\Lambda, we assume the outer expansions and the inner expansions hold simultaneously. Since Γε±\Gamma_{\varepsilon}^{\pm} are assumed to be graphs over Λ\Lambda, we introduce the functions Yε±(𝒔)Y_{\varepsilon}^{\pm}(\bm{s}) such that Γε±={z=Yε±(𝒔):𝒔U}\Gamma_{\varepsilon}^{\pm}=\{z=Y_{\varepsilon}^{\pm}(\bm{s})\,:\,\bm{s}\in U\}. Furthermore, we assume an expansion of the form

Yε±(𝒔)=Y0±(𝒔)+εY1±(𝒔)+𝒪(ε2)Y_{\varepsilon}^{\pm}(\bm{s})=Y_{0}^{\pm}(\bm{s})+\varepsilon Y_{1}^{\pm}(\bm{s})+\mathcal{O}(\varepsilon^{2})

is valid. As Γε±\Gamma_{\varepsilon}^{\pm} converge to Λ\Lambda, in the computations below we will deduce the values for Y0±(𝒔)Y_{0}^{\pm}(\bm{s}). Then, by comparing these two expansions in this intermediate region we infer matching conditions relating the outer expansions to the inner expansions via boundary conditions for the outer expansions. 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 [31] and [47, Appendix D])

B0(𝒔,z)\displaystyle B_{0}(\bm{s},z) {limδ0b0(𝒙+δ𝝂(𝒙))=:b0+(𝒙) for zY0+,limδ0b0(𝒙δ𝝂(𝒙))=:b0(𝒙) for zY0,\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 Y_{0}^{+},\\ \lim_{\delta\searrow 0}b_{0}(\bm{x}-\delta\bm{\nu}(\bm{x}))=:b_{0}^{-}(\bm{x})&\text{ for }z\to Y_{0}^{-},\end{cases}
zB0(𝒔,z)\displaystyle\partial_{z}B_{0}(\bm{s},z) 0 as zY0±,\displaystyle\to 0\text{ as }z\to Y_{0}^{\pm},
zB1(𝒔,z)\displaystyle\partial_{z}B_{1}(\bm{s},z) {limδ0(b0)(𝒙+δ𝝂(𝒙))𝝂(𝒙)=:b0+𝝂 for zY0+,limδ0(b0)(𝒙δ𝝂(𝒙))𝝂(𝒙)=:b0𝝂 for zY0,\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 Y_{0}^{+},\\ \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 Y_{0}^{-},\end{cases}

for 𝒙Λ\bm{x}\in\Lambda. Consequently, we denote the jump of a quantity bb across Λ\Lambda as

[b]+:=limδ0b(𝒙+δ𝝂(𝒙))limδ0b(𝒙δ𝝂(𝒙)) for 𝒙Λ.[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\Lambda.

5.2.4 Analysis of the inner expansions

We introduce the notation (𝑩)sym=12(𝑩+𝑩)\big{(}\bm{B}\big{)}^{\mathrm{sym}}=\frac{1}{2}(\bm{B}+\bm{B}^{\top}) for a second order tensor 𝑩\bm{B}. Plugging in the inner expansions to the state equation (2.3), to leading order 𝒪(1ε2)\mathcal{O}(\frac{1}{\varepsilon^{2}}) we find that

𝟎=𝝂z(𝒞(Φ0)(z𝑼0𝝂)sym).\bm{0}=\bm{\nu}\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}\Big{)}.

Taking the product with 𝑼0\bm{U}_{0} and by the symmetry of 𝒞\mathcal{C} we have the relation

𝝂z(𝒞(Φ0)(z𝑼0𝝂)sym)𝑼0=z(𝒞(Φ0)(z𝑼0𝝂)sym):𝑼0𝝂.\bm{\nu}\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}\Big{)}\cdot\bm{U}_{0}=\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}\Big{)}:\bm{U}_{0}\otimes\bm{\nu}.

Integrating over zz and by parts, using z𝝂=𝟎\partial_{z}\bm{\nu}=\bm{0} and by the matching condition limzY0±z𝑼0=𝟎\lim_{z\to Y_{0}^{\pm}}\partial_{z}\bm{U}_{0}=\bm{0} leads to

0=Y0Y0+𝒞(Φ0)(z𝑼0𝝂)sym:(z𝑼0𝝂)symdz.0=\int_{Y_{0}^{-}}^{Y_{0}^{+}}\mathcal{C}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}:\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}\,dz.

Coercivity of 𝒞\mathcal{C} yields that (z𝑼0𝝂)sym=𝟎\big{(}\partial_{z}\bm{U}_{0}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}=\bm{0} and hence z𝑼0(𝒔,z)=𝟎\partial_{z}\bm{U}_{0}(\bm{s},z)=\bm{0}. This implies 𝑼0\bm{U}_{0} is constant in zz and by the matching conditions we obtain

𝒖0+=𝑼0=𝒖0[𝒖0]+=𝟎.\displaystyle\bm{u}_{0}^{+}=\bm{U}_{0}=\bm{u}_{0}^{-}\quad\implies\quad[\bm{u}_{0}]_{-}^{+}=\bm{0}. (5.13)

Then, using z𝑼0=𝟎\partial_{z}\bm{U}_{0}=\bm{0}, to the next order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}) we obtain from the state equation (2.3)

𝟎=z(𝒞(Φ0)(z𝑼1𝝂+Λ𝑼0)sym𝝂).\displaystyle\bm{0}=\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{1}\otimes\bm{\nu}+\nabla_{\Lambda}\bm{U}_{0}\big{)}^{\mathrm{sym}}\bm{\nu}\Big{)}. (5.14)

Integrating over zz and using the matching condition

z𝑼1𝝂+Λ𝑼0{𝒖0+ for zY0+,𝒖0 for zY0,\partial_{z}\bm{U}_{1}\otimes\bm{\nu}+\nabla_{\Lambda}\bm{U}_{0}\to\begin{cases}\nabla\bm{u}_{0}^{+}&\text{ for }z\to Y_{0}^{+},\\ \nabla\bm{u}_{0}^{-}&\text{ for }z\to Y_{0}^{-},\end{cases}

we obtain

𝒞(1)(𝒖0+)𝝂𝒞(1)(𝒖0)𝝂=[𝒞(𝒖0)]+𝝂=𝟎.\displaystyle\mathcal{C}(1)\mathcal{E}(\bm{u}_{0}^{+})\bm{\nu}-\mathcal{C}(-1)\mathcal{E}(\bm{u}_{0}^{-})\bm{\nu}=[\mathcal{C}\mathcal{E}(\bm{u}_{0})]_{-}^{+}\bm{\nu}=\bm{0}. (5.15)

From the optimality condition (5.5a), to leading order 𝒪(1ε2)\mathcal{O}(\frac{1}{\varepsilon^{2}}) we obtain the trivial equality 0=00=0 due to z𝑼0=𝟎\partial_{z}\bm{U}_{0}=\bm{0}. Recalling (5.11) for μ1\mu_{-1}, to the next order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}) of (5.5a) we have

Ψ0(Φ0)|γ(𝝂)|2zzΦ0=0,\displaystyle\Psi_{0}^{\prime}(\Phi_{0})-|\gamma(\bm{\nu})|^{2}\partial_{zz}\Phi_{0}=0, (5.16)

where we again use z𝑼0=𝟎\partial_{z}\bm{U}_{0}=\bm{0} to see that no terms from the elastic part contribute at order 𝒪(1ε)\mathcal{O}(\frac{1}{\varepsilon}), the expression (5.12) for the divergence in the inner variables, the relation Dγ(𝝂)𝝂=γ(𝝂)\mathrm{D}\gamma(\bm{\nu})\cdot\bm{\nu}=\gamma(\bm{\nu}) from (5.7), the fact that γDγ\gamma\mathrm{D}\gamma is positively homogeneous of degree one, so that

(γDγ)(1εzΦ0𝝂)=1εzΦ0(γDγ)(𝝂),(\gamma\mathrm{D}\gamma)\big{(}\tfrac{1}{\varepsilon}\partial_{z}\Phi_{0}\bm{\nu}\big{)}=\tfrac{1}{\varepsilon}\partial_{z}\Phi_{0}(\gamma\mathrm{D}\gamma)(\bm{\nu}),

as well as the fact that 𝝂\bm{\nu} is independent of zz. We now construct a solution to (5.16) fulfilling the conditions Φ0(𝒔,z)±1\Phi_{0}(\bm{s},z)\to\pm 1 as zY0±(𝒔)z\to Y_{0}^{\pm}(\bm{s}) as per the matching conditions, and also satisfies zΦ0(𝒔,z)>0\partial_{z}\Phi_{0}(\bm{s},z)>0. Let ψ(z)\psi(z) be the monotone function defined in (5.3) which satisfies the boundary value problem

ψ′′(z)=Ψ0(ψ(z)),limz±π2ψ(z)=±1,\psi^{\prime\prime}(z)=\Psi_{0}^{\prime}(\psi(z)),\quad\lim_{z\to\pm\frac{\pi}{2}}\psi(z)=\pm 1,

with ψ(0)=0\psi(0)=0 and ψ(z)>0\psi^{\prime}(z)>0 for all zz\in\mathbb{R}. We define Φ0(𝒔,z)=ψ(zγ(𝝂(𝒔)))\Phi_{0}(\bm{s},z)=\psi\Big{(}\frac{z}{\gamma(\bm{\nu}(\bm{s}))}\Big{)}, and a short calculation shows that

zzΦ0(𝒔,z)=1|γ(𝝂(𝒔))|2ψ′′(zγ(𝝂(𝒔)))=1|γ(𝝂(𝒔))|2Ψ0(Φ0(𝒔,z)).\displaystyle\partial_{zz}\Phi_{0}(\bm{s},z)=\frac{1}{|\gamma(\bm{\nu}(\bm{s}))|^{2}}\psi^{\prime\prime}\Big{(}\frac{z}{\gamma(\bm{\nu}(\bm{s}))}\Big{)}=\frac{1}{|\gamma(\bm{\nu}(\bm{s}))|^{2}}\Psi_{0}^{\prime}(\Phi_{0}(\bm{s},z)).

The property ψ(±π2)=±1\psi(\pm\frac{\pi}{2})=\pm 1 and the matching conditions for Φ0\Phi_{0} provide the identifications

Y0+(𝒔)=π2γ(𝝂(𝒔)),Y0(𝒔)=π2γ(𝝂(𝒔)).Y_{0}^{+}(\bm{s})=\frac{\pi}{2}\gamma(\bm{\nu}(\bm{s})),\quad Y_{0}^{-}(\bm{s})=-\frac{\pi}{2}\gamma(\bm{\nu}(\bm{s})).

Moreover, since ψ\psi is monotone and γ(𝝂)\gamma(\bm{\nu}) is positive, we can infer that zΦ0(𝒔,z)>0\partial_{z}\Phi_{0}(\bm{s},z)>0. Now, multiplying (5.16) with zΦ0\partial_{z}\Phi_{0} and integrating over zz yields the so-called equipartition of energy

Ψ0(Φ0(𝒔,z))=|γ(𝝂(𝒔))|22|zΦ0(𝒔,z)|2,\Psi_{0}(\Phi_{0}(\bm{s},z))=\frac{|\gamma(\bm{\nu}(\bm{s}))|^{2}}{2}|\partial_{z}\Phi_{0}(\bm{s},z)|^{2},

after using Ψ0(1)=0\Psi_{0}(-1)=0 and limzY0zΦ0(𝒔,z)=0\lim_{z\to Y_{0}^{-}}\partial_{z}\Phi_{0}(\bm{s},z)=0. Performing another integration over zz leads to the relation

Y0Y0+|zΦ0|2dz=2|γ(𝝂(𝒔))|2Y0Y0+Ψ0(Φ0(𝒔,z))dz=112Ψ0(r)γ(𝝂(𝒔))dr=cΨγ(𝝂(𝒔)).\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}|\partial_{z}\Phi_{0}|^{2}\,\mathrm{d}z=\frac{2}{|\gamma(\bm{\nu}(\bm{s}))|^{2}}\int_{Y_{0}^{-}}^{Y_{0}^{+}}\Psi_{0}(\Phi_{0}(\bm{s},z))\,\mathrm{d}z=\int_{-1}^{1}\frac{\sqrt{2\Psi_{0}(r)}}{\gamma(\bm{\nu}(\bm{s}))}\,\mathrm{d}r=\frac{c_{\Psi}}{\gamma(\bm{\nu}(\bm{s}))}. (5.17)

To the next order 𝒪(1)\mathcal{O}(1), we find that

0\displaystyle 0 =2β𝕙(Φ0)𝒇𝑼0+μ0+α^Ψ0′′(Φ0)Φ1α^divΛ([γDγ](𝝂)zΦ0)\displaystyle=2\beta{\mathbbm{h}}^{\prime}(\Phi_{0})\bm{f}\cdot\bm{U}_{0}+\mu_{0}+\widehat{\alpha}\Psi_{0}^{\prime\prime}(\Phi_{0})\Phi_{1}-{\widehat{\alpha}}\mathrm{div}_{\Lambda}([\gamma\mathrm{D}\gamma](\bm{\nu})\partial_{z}\Phi_{0}) (5.18)
α^z(𝝂D(γDγ)(zΦ0𝝂)(zΦ1𝝂+ΛΦ0))\displaystyle\quad-{\widehat{\alpha}}\partial_{z}(\bm{\nu}\cdot\mathrm{D}(\gamma\mathrm{D}\gamma)(\partial_{z}\Phi_{0}\bm{\nu})(\partial_{z}\Phi_{1}\bm{\nu}+\nabla_{\Lambda}\Phi_{0}))
β𝒞(Φ0)(z𝑼1𝝂+Λ𝑼0)sym:(z𝑼1𝝂+Λ𝑼0)sym.\displaystyle\quad-\beta\mathcal{C}^{\prime}(\Phi_{0})\big{(}\partial_{z}\bm{U}_{1}\otimes\bm{\nu}+\nabla_{\Lambda}\bm{U}_{0}\big{)}^{\mathrm{sym}}:\big{(}\partial_{z}\bm{U}_{1}\otimes\bm{\nu}+\nabla_{\Lambda}\bm{U}_{0}\big{)}^{\mathrm{sym}}.

The aim is to multiply the above with zΦ0\partial_{z}\Phi_{0} and integrate over zz, but first, let us derive some useful relations. Using the symmetry 𝒞ijkl=𝒞jikl\mathcal{C}_{ijkl}=\mathcal{C}_{jikl} of the elasticity tensor and z𝑼0=𝟎\partial_{z}\bm{U}_{0}=\bm{0}, we deduce that, with the shorthand U=(z𝑼1𝝂+Λ𝑼0)sym\mathcal{E}_{U}=\big{(}\partial_{z}\bm{U}_{1}\otimes\bm{\nu}+\nabla_{\Lambda}\bm{U}_{0}\big{)}^{\mathrm{sym}},

𝒞(Φ0)U:zU=𝒞(Φ0)U:(zz𝑼1𝝂)sym=𝒞(Φ0)U𝝂zz𝑼1.\displaystyle\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\partial_{z}\mathcal{E}_{U}=\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\big{(}\partial_{zz}\bm{U}_{1}\otimes\bm{\nu}\big{)}^{\mathrm{sym}}=\mathcal{C}(\Phi_{0})\mathcal{E}_{U}\bm{\nu}\cdot\partial_{zz}\bm{U}_{1}.

Hence, together with z(𝒞(Φ0)U𝝂)=𝟎\partial_{z}(\mathcal{C}(\Phi_{0})\mathcal{E}_{U}\bm{\nu})=\bm{0} from (5.14) and the matching conditions we obtain the relation

Y0Y0+z𝒞(Φ0)U:Udz\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}\partial_{z}\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\mathcal{E}_{U}\,\mathrm{d}z =Y0Y0+z𝒞(Φ0)U:U+2𝒞(Φ0)U:zUdz\displaystyle=\int_{Y_{0}^{-}}^{Y_{0}^{+}}\partial_{z}\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\mathcal{E}_{U}+2\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\partial_{z}\mathcal{E}_{U}\,\mathrm{d}z
Y0Y0+2𝒞(Φ0)U𝝂zz𝑼1+2z(𝒞(Φ0)U𝝂)z𝑼1dz\displaystyle\quad-\int_{Y_{0}^{-}}^{Y_{0}^{+}}2\mathcal{C}(\Phi_{0})\mathcal{E}_{U}\bm{\nu}\cdot\partial_{zz}\bm{U}_{1}+2\partial_{z}(\mathcal{C}(\Phi_{0})\mathcal{E}_{U}\bm{\nu})\cdot\partial_{z}\bm{U}_{1}\,\mathrm{d}z
=Y0Y0+z(𝒞(Φ0)U:U)2z(𝒞(Φ0)U𝝂z𝑼1)dz\displaystyle=\int_{Y_{0}^{-}}^{Y_{0}^{+}}\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\mathcal{E}_{U}:\mathcal{E}_{U}\Big{)}-2\partial_{z}\Big{(}\mathcal{C}(\Phi_{0})\mathcal{E}_{U}\bm{\nu}\cdot\partial_{z}\bm{U}_{1}\Big{)}\,\mathrm{d}z
=[𝒞(𝒖0):(𝒖0)2𝒞(Φ0)(𝒖0)𝝂(𝒖0)𝝂]+.\displaystyle=[\mathcal{C}\mathcal{E}(\bm{u}_{0}):\mathcal{E}(\bm{u}_{0})-2\mathcal{C}(\Phi_{0})\mathcal{E}(\bm{u}_{0})\bm{\nu}\cdot(\nabla\bm{u}_{0})\bm{\nu}]_{-}^{+}.

By (5.8), (5.9), (5.16), (5.17) and the matching conditions of zΦ0\partial_{z}\Phi_{0}, we have

Y0Y0+Ψ0′′(Φ0)Φ1zΦ0dz=Y0Y0+Ψ0(Φ0)zΦ1dz=|γ(𝝂)|2Y0Y0+zzΦ0zΦ1dz,\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}\Psi_{0}^{\prime\prime}(\Phi_{0})\Phi_{1}\partial_{z}\Phi_{0}\,\mathrm{d}z=-\int_{Y_{0}^{-}}^{Y_{0}^{+}}\Psi_{0}^{\prime}(\Phi_{0})\partial_{z}\Phi_{1}\,\mathrm{d}z=-|\gamma(\bm{\nu})|^{2}\int_{Y_{0}^{-}}^{Y_{0}^{+}}\partial_{zz}\Phi_{0}\partial_{z}\Phi_{1}\,\mathrm{d}z,
Y0Y0+𝝂[D(γDγ)(𝝂)]𝝂(zzΦ1zΦ0)dz=|γ(𝝂)|2Y0Y0+zΦ1zzΦ0dz,\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}\bm{\nu}\cdot[\mathrm{D}(\gamma\mathrm{D}\gamma)(\bm{\nu})]\bm{\nu}(\partial_{zz}\Phi_{1}\partial_{z}\Phi_{0})\,\mathrm{d}z=-|\gamma(\bm{\nu})|^{2}\int_{Y_{0}^{-}}^{Y_{0}^{+}}\partial_{z}\Phi_{1}\partial_{zz}\Phi_{0}\,\mathrm{d}z,
Y0Y0+divΛ([γDγ](𝝂)zΦ0)zΦ0dz=cΨdivΛ([γDγ](𝝂))γ(𝝂)+[γDγ](𝝂)Λ(cΨ2γ(𝝂)),\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}\mathrm{div}_{\Lambda}([\gamma\mathrm{D}\gamma](\bm{\nu})\partial_{z}\Phi_{0})\partial_{z}\Phi_{0}\,\mathrm{d}z=\frac{c_{\Psi}\mathrm{div}_{\Lambda}([\gamma\mathrm{D}\gamma](\bm{\nu}))}{\gamma(\bm{\nu})}+[\gamma\mathrm{D}\gamma](\bm{\nu})\cdot\nabla_{\Lambda}\Big{(}\frac{c_{\Psi}}{2\gamma(\bm{\nu})}\Big{)},
Y0Y0+(𝝂[D(γDγ)(𝝂)]ΛzΦ0)zΦ0dz=[γDγ](𝝂)Λ(cΨ2γ(𝝂)).\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}(\bm{\nu}\cdot[\mathrm{D}(\gamma\mathrm{D}\gamma)(\bm{\nu})]\nabla_{\Lambda}\partial_{z}\Phi_{0})\partial_{z}\Phi_{0}\,\mathrm{d}z=[\gamma\mathrm{D}\gamma](\bm{\nu})\cdot\nabla_{\Lambda}\Big{(}\frac{c_{\Psi}}{2\gamma(\bm{\nu})}\Big{)}.

Hence,

Y0Y0+divΛ([γDγ](𝝂)zΦ0)zΦ0dz+Y0Y0+(𝝂[D(γDγ)(𝝂)]ΛzΦ0)zΦ0dz\displaystyle\int_{Y_{0}^{-}}^{Y_{0}^{+}}\mathrm{div}_{\Lambda}([\gamma\mathrm{D}\gamma](\bm{\nu})\partial_{z}\Phi_{0})\partial_{z}\Phi_{0}\,\mathrm{d}z+\int_{Y_{0}^{-}}^{Y_{0}^{+}}(\bm{\nu}\cdot[\mathrm{D}(\gamma\mathrm{D}\gamma)(\bm{\nu})]\nabla_{\Lambda}\partial_{z}\Phi_{0})\partial_{z}\Phi_{0}\,\mathrm{d}z
=cΨdivΛ([γDγ](𝝂))γ(𝝂)+[γDγ](𝝂)Λ(cΨγ(𝝂))=cΨdivΛ(Dγ(𝝂)),\displaystyle\quad=\frac{c_{\Psi}\mathrm{div}_{\Lambda}([\gamma\mathrm{D}\gamma](\bm{\nu}))}{\gamma(\bm{\nu})}+[\gamma\mathrm{D}\gamma](\bm{\nu})\cdot\nabla_{\Lambda}\Big{(}\frac{c_{\Psi}}{\gamma(\bm{\nu})}\Big{)}=c_{\Psi}\mathrm{div}_{\Lambda}\big{(}\mathrm{D}\gamma(\bm{\nu})\big{)},

and upon multiplying (5.18) with zΦ0\partial_{z}\Phi_{0} and integrate over zz, employ the matching conditions leads to the following solvability condition for Φ1\Phi_{1}:

0\displaystyle 0 =2β𝒇𝒖0[𝕙]++2μ0α^cΨdivΛ(Dγ(𝝂))\displaystyle=2\beta\bm{f}\cdot\bm{u}_{0}[{\mathbbm{h}}]_{-}^{+}+2\mu_{0}-\widehat{\alpha}c_{\Psi}\mathrm{div}_{\Lambda}(\mathrm{D}\gamma(\bm{\nu})) (5.19)
β[𝒞(𝒖0):(𝒖0)2𝒞(𝒖0)𝝂(𝒖0)𝝂]+ on Λ.\displaystyle\quad-\beta[\mathcal{C}\mathcal{E}(\bm{u}_{0}):\mathcal{E}(\bm{u}_{0})-2\mathcal{C}\mathcal{E}(\bm{u}_{0})\bm{\nu}\cdot(\nabla\bm{u}_{0})\bm{\nu}]_{-}^{+}\quad\text{ on }\Lambda.

Note that the assumed higher regularity 𝒇H1(Ω,d)\bm{f}\in H^{1}(\Omega,\mathbb{R}^{d}) allows us to define 𝒇\bm{f} on Λ\Lambda. Thus, the sharp interface limit ε0\varepsilon\to 0 consists of the system (5.10) posed in Ω±\Omega_{\pm}, along with the boundary conditions (5.13), (5.15) and (5.19) on Λ\Lambda.

Remark 5.1 (Isotropic case).

In the isotropic case γ(𝛎)=|𝛎|\gamma(\bm{\nu})=|\bm{\nu}|, for unit normals 𝛎\bm{\nu} of Λ\Lambda, we have that Dγ(𝛎)=𝛎|𝛎|=𝛎\mathrm{D}\gamma(\bm{\nu})=\frac{\bm{\nu}}{|\bm{\nu}|}=\bm{\nu} and divΛ(Dγ(𝛎))=κΛ\mathrm{div}_{\Lambda}(\mathrm{D}\gamma(\bm{\nu}))=-\kappa_{\Lambda}, where κΛ\kappa_{\Lambda} is the mean curvature of Λ\Lambda. Then, choosing 𝕙(φ)=1{\mathbbm{h}}(\varphi)=1 so that [𝕙]+=0[{\mathbbm{h}}]_{-}^{+}=0, we observe that (5.19) simplifies to

0=2μ0+α^cΨκΛβ[𝒞(𝒖0):(𝒖0)2𝒞(𝒖0)𝝂(𝒖0)𝝂)]+ on Λ,0=2\mu_{0}+\widehat{\alpha}c_{\Psi}\kappa_{\Lambda}-\beta[\mathcal{C}\mathcal{E}(\bm{u}_{0}):\mathcal{E}(\bm{u}_{0})-2\mathcal{C}\mathcal{E}(\bm{u}_{0})\bm{\nu}\cdot(\nabla\bm{u}_{0})\bm{\nu})]_{-}^{+}\text{ on }\Lambda,

which coincides with the formula obtained from [23, (24)] with adjoint 𝐪0=β𝐮0\bm{q}_{0}=\beta\bm{u}_{0}, zero eigenstrain ¯=𝟎\overline{\mathcal{E}}=\bm{0}, Lagrange multiplier λ0=μ0\lambda_{0}=\mu_{0} for the mass constraint, and hΩ(𝐱,𝐮)=β𝐟𝐮h_{\Omega}(\bm{x},\bm{u})=\beta\bm{f}\cdot\bm{u} that satisfies [hΩ(𝐱,𝐮0)]+=0[h_{\Omega}(\bm{x},\bm{u}_{0})]_{-}^{+}=0 thanks to (5.13).

5.3 Relation to shape derivatives derived in [4]

To relate our work with the setting of [4], we first set 𝒇=𝟎\bm{f}=\bm{0} in (5.10), and neglect all the equations posed in Ω\Omega_{-}. Then, the state equation [4, (2.1)] can be obtained by writing 𝒞(+1)\mathcal{C}(+1) as AA, Ω+\Omega_{+} as Ω\Omega in (5.10). Note that in this setting the free boundary Λ\Lambda is identified with the traction-free part Γ0Ω+\Gamma_{0}\cap\partial\Omega_{+} of Ω+\partial\Omega_{+}. Furthermore, the normal vector 𝒏\bm{n} in [4] on Λ\Lambda is equal to 𝝂-\bm{\nu} in our notation.

Writing φ(𝒏)\varphi(\bm{n}) as γ(𝝂)\gamma(-\bm{\nu}), the anisotropic perimeter functional PgP_{g} defined in [4, (3.1)] can be expressed as in (3.2), and in terms of our notation its shape derivative is given as (compare [4, Prop. 3.1])

Pg(Ω+)[𝜽]=ΛκΛγ(𝝂)𝜽𝝂+Dγ(𝝂)Λ(𝜽𝝂)dd1,P_{g}^{\prime}(\Omega_{+})[\bm{\theta}]=\int_{\Lambda}-\kappa_{\Lambda}\gamma(\bm{\nu})\bm{\theta}\cdot\bm{\nu}+\mathrm{D}\gamma(\bm{\nu})\cdot\nabla_{\Lambda}(\bm{\theta}\cdot\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1},

for admissible vector fields 𝜽\bm{\theta} of sufficient smoothness, where we used that Dγ\mathrm{D}\gamma is positively homogeneous of degree zero (cf. (5.7)) and the identity κΛ=divΛ𝝂=divΛ𝒏\kappa_{\Lambda}=-\mathrm{div}_{\Lambda}\bm{\nu}=\mathrm{div}_{\Lambda}\bm{n}. Here we point out a typo in the manuscript, where Ω(φ(𝒏))\nabla_{\partial\Omega}(\varphi(\bm{n})) written there should in fact be Dφ(𝒏)\mathrm{D}\varphi(\bm{n}) (compare [34, Prop. 8]).

Now, consider the shape optimization problem of minimizing L(Ω+):=βJ(Ω+)+α^cΨPg(Ω+)L(\Omega_{+}):=\beta J(\Omega_{+})+{\widehat{\alpha}}c_{\Psi}P_{g}(\Omega_{+}) with the mean compliance

J(Ω+):=Ω+𝒞(1)(𝒖):(𝒖)dx.J(\Omega_{+}):=\int_{\Omega_{+}}\mathcal{C}(1)\mathcal{E}(\bm{u}):\mathcal{E}(\bm{u})\,\mathrm{dx}.

Then, with sufficiently smooth solutions, its shape derivative in our notation is (see [4, p. 299])

L(Ω+)[𝜽]\displaystyle L^{\prime}(\Omega_{+})[\bm{\theta}] =Λ[β𝒞(1)(𝒖):(𝒖)α^cΨκΛγ(𝝂)]𝜽𝝂+α^cΨDγ(𝝂)Λ(𝜽𝝂)dd1\displaystyle=\int_{\Lambda}[-\beta\mathcal{C}(1)\mathcal{E}(\bm{u}):\mathcal{E}(\bm{u})-\widehat{\alpha}c_{\Psi}\kappa_{\Lambda}\gamma(\bm{\nu})]\bm{\theta}\cdot\bm{\nu}+\widehat{\alpha}c_{\Psi}\mathrm{D}\gamma(\bm{\nu})\cdot\nabla_{\Lambda}(\bm{\theta}\cdot\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1}
=Λ[β𝒞(1)(𝒖):(𝒖)α^cΨdivΛ(Dγ(𝝂))]𝜽𝝂dd1,\displaystyle=\int_{\Lambda}[-\beta\mathcal{C}(1)\mathcal{E}(\bm{u}):\mathcal{E}(\bm{u})-\widehat{\alpha}c_{\Psi}\mathrm{div}_{\Lambda}(\mathrm{D}\gamma(\bm{\nu}))]\bm{\theta}\cdot\bm{\nu}\,\mathrm{d}\mathcal{H}^{d-1},

where in the above we have applied the integration by parts formula

ΛΛ(𝜽𝝂)Dγ(𝝂)dd1=Λ[κΛDγ(𝝂)𝝂divΛ(Dγ(𝝂))](𝜽𝝂)dd1,\int_{\Lambda}\nabla_{\Lambda}(\bm{\theta}\cdot\bm{\nu})\cdot\mathrm{D}\gamma(\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1}=\int_{\Lambda}[\kappa_{\Lambda}\mathrm{D}\gamma(\bm{\nu})\cdot\bm{\nu}-\mathrm{div}_{\Lambda}(\mathrm{D}\gamma(\bm{\nu}))](\bm{\theta}\cdot\bm{\nu})\,\mathrm{d}\mathcal{H}^{d-1},

as well as the relation Dγ(𝝂)𝝂=γ(𝝂)\mathrm{D}\gamma(\bm{\nu})\cdot\bm{\nu}=\gamma(\bm{\nu}) from (5.7) to cancel the terms involving κΛ\kappa_{\Lambda}. Hence, the strong formulation of L(Ω+)[𝜽]=0L^{\prime}(\Omega_{+})[\bm{\theta}]=0 is

0=α^cΨdivΛ(Dγ(𝝂))+β𝒞(1)(𝒖):(𝒖) on Λ=Γ0Ω+.0=\widehat{\alpha}c_{\Psi}\mathrm{div}_{\Lambda}(\mathrm{D}\gamma(\bm{\nu}))+\beta\mathcal{C}(1)\mathcal{E}(\bm{u}):\mathcal{E}(\bm{u})\quad\text{ on }\Lambda=\Gamma_{0}\cap\partial\Omega_{+}.

This coincides with the identity obtained if we neglect the Lagrange multiplier, set 𝒞(1)=𝟎\mathcal{C}(-1)=\bm{0} and 𝒇=𝟎\bm{f}=\bm{0} in (5.19), and also use the relation 𝒞(+1)(𝒖0+)𝝂=𝟎\mathcal{C}(+1)\mathcal{E}(\bm{u}_{0}^{+})\bm{\nu}=\bm{0} from (5.15).

6 Numerical approximation

6.1 BGN anisotropies

For the numerical approximation of an optimal design variable φ¯𝒱m\overline{\varphi}\in\mathcal{V}_{m} to the structural optimization problem (P), we first consider a discretization of the variational inequality (5.4) with differentiable γ\gamma. In general, the term γDγ\gamma\mathrm{D}\gamma is highly nonlinear, and to facilitate the subsequent discussion regarding its discretization we first consider a matrix-type reformulation of the anisotropy density function introduced by the works of the first and third authors, see, e.g., [14, 15, 16] for more details.

Suppose for some LL\in\mathbb{N}, the anisotropy density function γ\gamma can be expressed as

γ(𝒒)=l=1Lγl(𝒒), where γl(𝒒):=𝒒𝐆l𝒒 for 𝒒d,\displaystyle\gamma(\bm{q})=\sum_{l=1}^{L}\gamma_{l}(\bm{q}),\text{ where }\gamma_{l}(\bm{q}):=\sqrt{\bm{q}\cdot{\bf{G}}_{l}\bm{q}}\text{ for }\bm{q}\in\mathbb{R}^{d}, (6.1)

with symmetric positive definite matrices 𝐆ld×d{\bf{G}}_{l}\in\mathbb{R}^{d\times d} for l=1,,Ll=1,\dots,L. Then, a short calculation shows that

Dγ(𝒒)=l=1L𝐆l𝒒γl(𝒒)𝒒d{𝟎},\mathrm{D}\gamma(\bm{q})=\sum_{l=1}^{L}\frac{{\bf{G}}_{l}\bm{q}}{\gamma_{l}(\bm{q})}\quad\forall\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\},

and recalling that A():=12|γ()|2A(\cdot):=\frac{1}{2}|\gamma(\cdot)|^{2}, we infer also

DA(𝒒)=[γDγ](𝒒)=γ(𝒒)l=1L𝐆l𝒒γl(𝒒)𝒒d{𝟎}.\mathrm{D}A(\bm{q})=[\gamma\mathrm{D}\gamma](\bm{q})=\gamma(\bm{q})\sum_{l=1}^{L}\frac{{\bf{G}}_{l}\bm{q}}{\gamma_{l}(\bm{q})}\quad\forall\bm{q}\in\mathbb{R}^{d}\setminus\{\bm{0}\}.

For 𝒑\bm{p} close to 𝒒\bm{q}, we now perform a linearization DA(𝒒)𝐁(𝒑)𝒒\mathrm{D}A(\bm{q})\approx{\bf{B}}(\bm{p})\bm{q}, where

𝐁(𝒑):={γ(𝒑)l=1L𝐆lγl(𝒑) if 𝒑𝟎,Ll=1L𝐆l if 𝒑=𝟎,\displaystyle{\bf{B}}(\bm{p}):=\begin{cases}\displaystyle\gamma(\bm{p})\sum_{l=1}^{L}\frac{{\bf{G}}_{l}}{\gamma_{l}(\bm{p})}&\text{ if }\bm{p}\neq\bm{0},\\[10.0pt] \displaystyle L\sum_{l=1}^{L}{\bf{G}}_{l}&\text{ if }\bm{p}=\bm{0},\end{cases}

so that

𝐁(𝒒)𝒒=DA(𝒒) for all 𝒒d.{\bf{B}}(\bm{q})\bm{q}=\mathrm{D}A(\bm{q})\text{ for all }\bm{q}\in\mathbb{R}^{d}.

Furthermore, as {𝐆l}l=1L\{{\bf{G}}_{l}\}_{l=1}^{L} are symmetric positive definite, the matrix 𝐁(𝒑){\bf{B}}(\bm{p}) is also symmetric positive definite for all 𝒑d\bm{p}\in\mathbb{R}^{d}. We term anisotropies γ\gamma that are of the form (6.1) as BGN anisotropies. Notice that for the definition of the linearized matrix 𝐁{\bf{B}} it is sufficient to have anisotropy densities γC0(d)\gamma\in C^{0}(\mathbb{R}^{d}). However, due to our examples (3.10) and (3.11) it will turn out that the corresponding 𝐆l{\bf{G}}_{l} are no longer constant matrices. Thus, inspired by the works [14, 15, 16] we extend their approach to the case where 𝐆l{\bf{G}}_{l} are piecewise constant and dependent on 𝒒\bm{q}, as shown in the following examples. In our numerical approximation to the optimality condition detailed in the next section we simply replace DA(𝒒)=[γDγ](𝒒)\mathrm{D}A(\bm{q})=[\gamma\mathrm{D}\gamma](\bm{q}) by 𝐁(𝒒)𝒒{\bf{B}}(\bm{q})\bm{q}, where 𝐁(𝒒){\bf{B}}(\bm{q}) now involves piecewise constant matrices 𝐆l(𝒒){\bf{G}}_{l}(\bm{q}).

Example 6.1.

The convex anisotropy (3.10) can be expressed as an extended BGN anisotropy

γα(𝒒)=𝒒𝐆(𝒒)𝒒,\displaystyle\gamma_{\alpha}(\bm{q})=\sqrt{\bm{q}\cdot{\bf{G}}(\bm{q})\bm{q}},

where for 𝐪=(q1,,qd)d\bm{q}=(q_{1},\dots,q_{d})\in\mathbb{R}^{d},

𝐆(𝒒)={𝐈 if qdα|𝒒|,diag(0,0,,0,α2) if qd<α|𝒒|.\displaystyle{\bf{G}}(\bm{q})=\begin{cases}\displaystyle{\bf I}&\displaystyle\text{ if }q_{d}\geq-\alpha|\bm{q}|,\\[10.0pt] \displaystyle\mathrm{diag}\big{(}0,0,\dots,0,\alpha^{-2}\big{)}&\displaystyle\text{ if }q_{d}<-\alpha|\bm{q}|.\end{cases}

In turn, this provides us with the linearization matrix 𝐁(𝐪)=𝐆(𝐪){\bf{B}}(\bm{q})={\bf{G}}(\bm{q}).

Example 6.2.

The nonconvex anisotropy (3.11) can be written as an extended BGN anisotropy

γα,λ(𝒒)=𝒒𝐆(𝒒)𝒒,\displaystyle\gamma_{\alpha,\lambda}(\bm{q})=\sqrt{\bm{q}\cdot{\bf{G}}(\bm{q})\bm{q}},

where for 𝐪2\bm{q}\in\mathbb{R}^{2},

𝐆(𝒒)={𝐈 if q2α|𝒒|,((1λλ)211α21λλ2α11α21λλ2α11α21λ2α2) if q10,q2<α|𝒒|,((1λλ)211α21λλ2α11α21λλ2α11α21λ2α2) if q1>0,q2<α|𝒒|.\displaystyle{\bf{G}}(\bm{q})=\begin{cases}\displaystyle{\bf I}&\displaystyle\text{ if }q_{2}\geq-\alpha|\bm{q}|,\\[10.0pt] \displaystyle\begin{pmatrix}\big{(}\frac{1-\lambda}{\lambda}\big{)}^{2}\frac{1}{1-\alpha^{2}}&-\frac{1-\lambda}{\lambda^{2}\alpha}\frac{1}{\sqrt{1-\alpha^{2}}}\\ -\frac{1-\lambda}{\lambda^{2}\alpha}\frac{1}{\sqrt{1-\alpha^{2}}}&\frac{1}{\lambda^{2}\alpha^{2}}\end{pmatrix}&\displaystyle\text{ if }q_{1}\leq 0,\,q_{2}<-\alpha|\bm{q}|,\\[15.0pt] \displaystyle\begin{pmatrix}\big{(}\frac{1-\lambda}{\lambda}\big{)}^{2}\frac{1}{1-\alpha^{2}}&\frac{1-\lambda}{\lambda^{2}\alpha}\frac{1}{\sqrt{1-\alpha^{2}}}\\ \frac{1-\lambda}{\lambda^{2}\alpha}\frac{1}{\sqrt{1-\alpha^{2}}}&\frac{1}{\lambda^{2}\alpha^{2}}\end{pmatrix}&\displaystyle\text{ if }q_{1}>0,\,q_{2}<-\alpha|\bm{q}|.\end{cases}

Then, the linearized matrix is given by 𝐁(𝐪)=𝐆(𝐪){\bf{B}}(\bm{q})={\bf{G}}(\bm{q}).

For numerical simulations we also consider regularizations of the convex anisotropy γα\gamma_{\alpha} in (3.10) of the form

γ^α,δ(𝒒):=γα(𝒒)+δ|𝒒|\displaystyle\widehat{\gamma}_{\alpha,\delta}(\bm{q}):=\gamma_{\alpha}(\bm{q})+\delta|\bm{q}| (6.2)

for δ>0\delta>0. This can also be expressed as an extended BGN anisotropy with L=2L=2, where we take γ1(𝒒)=γα(𝒒)\gamma_{1}(\bm{q})=\gamma_{\alpha}(\bm{q}), γ2(𝒒)=δ|𝒒|\gamma_{2}(\bm{q})=\delta|\bm{q}|,

𝐆1(𝒒)={𝐈 if qdα|𝒒|,diag(0,0,,0,α2) if qd<α|𝒒|, and 𝐆2(𝒒)=δ2𝐈.{\bf{G}}_{1}(\bm{q})=\begin{cases}\displaystyle{\bf I}&\displaystyle\text{ if }q_{d}\geq-\alpha|\bm{q}|,\\[10.0pt] \displaystyle\mathrm{diag}\big{(}0,0,\dots,0,\alpha^{-2}\big{)}&\displaystyle\text{ if }q_{d}<-\alpha|\bm{q}|,\end{cases}\quad\text{ and }\quad{\bf{G}}_{2}(\bm{q})=\delta^{2}{\bf I}{.}

The linearization matrix 𝐁(𝒒){\bf{B}}(\bm{q}) has the form

𝐁(𝒒)={γ^α,δ(𝒒)([γα(𝒒)]1𝐆1(𝒒)+|𝒒|1δ𝐈) if 𝒒𝟎,2(1+δ2)𝐈 if 𝒒=𝟎.{\bf{B}}(\bm{q})=\begin{cases}\widehat{\gamma}_{\alpha,\delta}(\bm{q})\Big{(}[\gamma_{\alpha}(\bm{q})]^{-1}{\bf{G}}_{1}(\bm{q})+|\bm{q}|^{-1}\delta{\bf I}\Big{)}&\text{ if }\bm{q}\neq\bm{0},\\ 2(1+\delta^{2}){\bf I}&\text{ if }\bm{q}=\bm{0}.\end{cases}

In Figure 7 we visualize the Frank diagram and Wulff shape of the anisotropy (6.2) in two spatial dimensions with δ=0.1\delta=0.1 and α=0.9\alpha=0.9, 0.70.7, 0.50.5, 0.30.3.

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 7: The Frank diagram (left column) and Wulff shape (right column) for the regularized anisotropy (6.2) with δ=0.1\delta=0.1. From top to bottom: α=0.9\alpha=0.9, 0.70.7, 0.50.5, 0.30.3. Notice the bottom of each Frank diagram is not exactly a straight line, but slightly curved.

6.2 Finite element discretization

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={χC0(Ω¯):χ|oP1(o)o𝒯h}and𝑺h=Sh××Sh=[Sh]d,\displaystyle S^{h}=\left\{\chi\in C^{0}(\overline{\Omega}):\,\chi_{|_{o}}\in P_{1}(o)\,\forall o\in\mathcal{T}_{h}\right\}\quad\text{and}\quad\bm{S}^{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. In addition we define the convex subsets

𝒱h={χSh:|χ|1 in Ω¯}and𝒱mh={χ𝒱h:(χ,1)=m(1,1)},\mathcal{V}^{h}=\left\{\chi\in S^{h}:|\chi|\leq 1\text{ in }\overline{\Omega}\right\}\quad\text{and}\quad\mathcal{V}^{h}_{m}=\left\{\chi\in{\mathcal{V}}^{h}:(\chi,1)=m(1,1)\right\}, (6.3)

where (,)(\cdot,\cdot) denotes the L2L^{2}–inner product on Ω\Omega, as well as

𝑺Dh={𝜼𝑺h:𝜼=𝟎 on ΓD}.\bm{S}^{h}_{D}=\left\{\bm{\eta}\in\bm{S}^{h}:\bm{\eta}={\bf 0}\text{ on }\Gamma_{D}\right\}. (6.4)

We now introduce a finite element approximation of the structural optimization problem (P) and the optimality conditions described above. Given φhn1𝒱mh\varphi_{h}^{n-1}\in\mathcal{V}^{h}_{m}, find (𝒖hn,φhn)𝑺Dh×𝒱mh(\bm{u}_{h}^{n},\varphi_{h}^{n})\in\bm{S}^{h}_{D}\times\mathcal{V}^{h}_{m} such that

(𝒖hn),(𝜼)𝒞(φhn1)=(𝕙(φhn1)𝒇,𝜼)h+Γg𝒈𝜼𝜼𝑺Dh,\displaystyle\langle\mathcal{E}(\bm{u}_{h}^{n}),\mathcal{E}(\bm{\eta})\rangle_{\mathcal{C}(\varphi_{h}^{n-1})}=\left({\mathbbm{h}}(\varphi_{h}^{n-1})\bm{f},\bm{\eta}\right)^{h}+\int_{\Gamma_{g}}\bm{g}\cdot\bm{\eta}\qquad\forall\bm{\eta}\in\bm{S}^{h}_{D}\,, (6.5a)
(ετ(φhnφhn1)α^εφhn,χφhn)h+α^ε(𝐁(φhn1)φhn,(χφhn))\displaystyle\left(\tfrac{\varepsilon}{\tau}(\varphi_{h}^{n}-\varphi_{h}^{n-1})-\tfrac{\widehat{\alpha}}{\varepsilon}\varphi_{h}^{n},\chi-\varphi_{h}^{n}\right)^{h}+\widehat{\alpha}\varepsilon({\bf{B}}(\varphi_{h}^{n-1})\nabla\varphi_{h}^{n},\nabla(\chi-\varphi_{h}^{n}))
(𝒖hn),(𝒖hn)(χφhn)𝒞(φhn1)χ𝒱mh.\displaystyle\qquad\geq\langle\mathcal{E}(\bm{u}_{h}^{n}),\mathcal{E}(\bm{u}_{h}^{n})(\chi-\varphi_{h}^{n})\rangle_{\mathcal{C}^{\prime}(\varphi_{h}^{n-1})}\qquad\forall\chi\in\mathcal{V}^{h}_{m}\,. (6.5b)

Here τ\tau denotes the time step size, (,)h(\cdot,\cdot)^{h} is the usual mass lumped L2L^{2}–inner product on Ω\Omega, and 𝑨,𝑩𝒞=Ω𝒞𝑨:𝑩dx\langle\bm{A},\bm{B}\rangle_{\mathcal{C}}=\int_{\Omega}\mathcal{C}\bm{A}:\bm{B}\,\mathrm{dx} for any fourth order tensor 𝒞\mathcal{C} and any matrices 𝑨\bm{A} and 𝑩\bm{B}. We implemented the scheme (6.5) with the help of the finite element toolbox ALBERTA, see [77]. 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 [11, 18] for a more detailed description. Clearly, the system (6.5a) decouples, and so we first solve the linear system (6.5a) in order to obtain 𝒖hn\bm{u}_{h}^{n}, and then solve the variational inequality (6.5b) for φhn\varphi_{h}^{n}. We employ the package LDL, see [33], together with the sparse matrix ordering AMD, see [9], in order to solve (6.5a). For the variational inequality (6.5b) we employ a secant method as described in [26] to satisfy the mass constraint (φhn,1)=m(\varphi_{h}^{n},1)=m, and use a nonlinear multigrid method similar to [55] for solving the variational inequalities over 𝒱h{\mathcal{V}}^{h} that arise as the inner problems for the secant iterations. The second method always converged in at most five steps. Finally, to increase the efficiency of the numerical computations in this paper, at times we exploit the symmetry of the problem and perform the computations in question only on half of the desired domain Ω\Omega. In those cases we prescribe “free-slip” boundary conditions for the displacement field 𝒖hn\bm{u}_{h}^{n} on the symmetry plane ΓDi\Gamma_{D_{i}}, that is, we replace (6.4) with

𝑺Dh={𝜼=(η1,,ηd)𝑺h:𝜼=𝟎 on ΓD and ηi=0 on ΓDi,i=1,,d}.\bm{S}^{h}_{D}=\left\{\bm{\eta}=(\eta_{1},\ldots,\eta_{d})\in\bm{S}^{h}:\bm{\eta}={\bf 0}\text{ on }\Gamma_{D}\ \text{ and }\ \eta_{i}=0\text{ on }\Gamma_{D_{i}},i=1,\ldots,d\right\}. (6.6)

All computations performed in this work are for spatial dimension d=2d=2. For the remainder of the paper we consider the quadratic interpolation function (2.2) in the elasticity tensor 𝒞(φ)\mathcal{C}(\varphi), forcing 𝒇=𝟎\bm{f}=\bm{0}, objective functional weightings β=1\beta=1 and α^=1\widehat{\alpha}=1 unless further specified.

6.3 Numerical simulations

6.3.1 Optimality condition without elasticity

We first investigate the setting with 𝒈=𝟎\bm{g}=\bm{0}, so that (6.5a) yields 𝒖hn=𝟎\bm{u}_{h}^{n}=\bm{0} and (6.5b) reduces to an Allen–Cahn variational inequality on 𝒱mh\mathcal{V}_{m}^{h}. From the discussion in Section 3, this is a phase field approximation of the volume preserving anisotropic mean curvature flow (due to the mass constraint), and it is expected that the long time behavior of solutions to display the Wulff shape (3.6) corresponding to the anisotropy γ\gamma, see Figure 7. In Figure 8 we display snapshots of the solution at times t=0t=0, 0.0010.001, 0.0050.005 and 0.030.03 with anisotropy (6.2) for parameter values α=0.5\alpha=0.5, δ=0.1\delta=0.1 and ε=1/(16π)\varepsilon=1/(16\pi). We notice from the bottom plot of the discrete Ginzburg–Landau energy γh(φhn)=(εA(φhn)+1εΨ(φhn),1)h\mathcal{E}_{\gamma}^{h}(\varphi_{h}^{n})=(\varepsilon A(\nabla\varphi_{h}^{n})+\tfrac{1}{\varepsilon}\Psi(\varphi_{h}^{n}),1)^{h} that it is decreasing over time, and on the top right we observe the expected Wulff shape attained near equilibrium.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: (Top) Snapshots of the solution with anisotropy (6.2) for α=0.5\alpha=0.5, δ=0.1\delta=0.1 and ε=1/(16π)\varepsilon=1/(16\pi) at times t=0t=0, 0.0010.001, 0.0050.005, 0.030.03. (Below) A plot of discrete Ginzburg–Landau energy γh(φhn)=(εA(φhn)+1εΨ(φhn),1)h\mathcal{E}_{\gamma}^{h}(\varphi_{h}^{n})=(\varepsilon A(\nabla\varphi_{h}^{n})+\tfrac{1}{\varepsilon}\Psi(\varphi_{h}^{n}),1)^{h} over time.

6.3.2 Dripping effect of a straight interface

Next, we investigate the role of the convexity/non-convexity of γ\gamma in the generation of the dripping effect [4, 10, 75]. We solve (6.5) with 𝒈=𝟎\bm{g}=\bm{0} and initial condition taken as some large perturbation of a straight interface. Taking (6.2) as the anisotropy with α=0.5\alpha=0.5, δ=0.1\delta=0.1 and ε=1/(32π)\varepsilon=1/(32\pi), in Figure 9 we display the snapshots of the solution at times t=0t=0, 0.0010.001, 0.0050.005 and 0.020.02, where it is clear that the oscillations are dampened over time. Thus, with a convex anisotropy the dripping effect is suppressed. In contrast, taking γ\gamma as the non-convex anisotropy (3.11) with λ=0.5\lambda=0.5 now yields Figure 10, where the corresponding snapshots of the solution at times t=0t=0, 0.0010.001, 0.0050.005, 0.020.02 are displayed. Here we clearly observe the behavior described in Example 3.3.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Snapshots of the solution with convex anisotropy (6.2) for α=0.5\alpha=0.5, δ=0.1\delta=0.1 and ε=1/(32π)\varepsilon=1/(32\pi) at times t=0t=0, 0.0010.001, 0.0050.005, 0.020.02.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Snapshots of the solution with non-convex anisotropy (3.11) for α=0.5\alpha=0.5, δ=0.1\delta=0.1, λ=0.5\lambda=0.5 and ε=1/(32π)\varepsilon=1/(32\pi) at times t=0t=0, 0.0010.001, 0.0050.005, 0.020.02.

6.3.3 Cantilever beam computation

Inspired by [4, Figure 8], for the state system (2.3), we consider the setting

Ω\displaystyle\Omega =(12,12)×(12,12),ΓD=(14,14)×{12},\displaystyle=(-\tfrac{1}{2},\tfrac{1}{2})\times(-\tfrac{1}{2},\tfrac{1}{2}),\quad\Gamma_{D}=(-\tfrac{1}{4},\tfrac{1}{4})\times\{-\tfrac{1}{2}\}, (6.7)
Γg\displaystyle\Gamma_{g} =(0.02,0.02)×{12},𝒈=(g0),\displaystyle=(-0.02,0.02)\times\{\tfrac{1}{2}\},\quad\bm{g}=\begin{pmatrix}g\\ 0\end{pmatrix},

with Γ0=Ω(ΓDΓg)\Gamma_{0}=\partial\Omega\setminus(\Gamma_{D}\cup\Gamma_{g}) and gg a positive constant. Note that the domain Ω\Omega we use here is larger than the setting in [4]. This is done to to avoid the influence of the domain boundary Ω\partial\Omega on the growth of the interfacial layer {|φh|<1}\{|\varphi_{h}|<1\}. As in [4] we consider an elastic material with a normalized Young’s modulus E=1E=1 and a Poisson ratio ν=0.33\nu=0.33. The corresponding Lamé constants in the elasticity tensor (2.1) can be calculated through the well-known relations

μ=E2(1+ν)andλ=Eν(1+ν)(12ν).\mu=\frac{E}{2(1+\nu)}\quad\text{and}\quad\lambda=\frac{E\nu}{(1+\nu)(1-2\nu)}. (6.8)

For g=5g=5 and α^=0.5\widehat{\alpha}=0.5, we solve the full optimality condition (6.5) with the convex regularized anisotropy (6.2) and compare the steady states for δ=0.1\delta=0.1 and a range of α=1,0.7,0.5,0.2\alpha{=1,0.7,0.5,0.2}. Note that smaller values of α\alpha indicate a stronger anisotropy. The results are displayed in Figure 11, where we taken random initial data φh0\varphi_{h}^{0} with mass m=1|Ω|Ωφh0dx=0.7m=\frac{1}{|\Omega|}\int_{\Omega}\varphi_{h}^{0}\,\mathrm{dx}=0.7. Note that α=1\alpha=1 corresponds to the isotropic case, see Remark 3.2, and as α\alpha decreases, we observe the interior connecting bridges in the cantilever beam become steeper in slope, up until α=0.2\alpha=0.2 where a design without interior structure is favored.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: Cantilever beam computation for ε=1/(32π)\varepsilon=1/(32\pi) and 𝒈=(5,0)\bm{g}=(5,0)^{\top} on [0.02,0.02]×{0.5}[-0.02,0.02]\times\{0.5\}. The solutions at time t=0.04t=0.04, starting from random initial data with mass m=0.7m=0.7. Anisotropy (6.2) used with δ=0.1\delta=0.1. From left to right: α=1\alpha=1 (isotropic case), 0.70.7, 0.50.5, 0.20.2.
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 12: Cantilever beam computation for ε=1/(32π)\varepsilon=1/(32\pi) and 𝒈=(g,0)\bm{g}=(g,0)^{\top} on [0.02,0.02]×{0.5}[-0.02,0.02]\times\{0.5\}. The solutions at time t=0.04t=0.04, starting from random initial data with mass m=0.7m=0.7. Anisotropy (6.2) used with δ=0.1\delta=0.1. (Top) g=30g=30 with α=1\alpha=1, 0.30.3, 0.20.2, 0.10.1. (Bottom) g=50g=50 with α=1\alpha=1, 0.30.3, 0.20.2, 0.10.1. Left-most column corresponds to the isotropic case.
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 13: Cantilever beam computation for ε=1/(64π)\varepsilon=1/(64\pi) and with 𝒈=(g,0)\bm{g}=(g,0)^{\top} on [0.02,0.02]×{0.5}[-0.02,0.02]\times\{0.5\}. The solutions at time t=0.01t=0.01, starting from the final states displayed in Figure 12. Anisotropy (6.2) used with δ=0.1\delta=0.1. (Top) g=30g=30 with α=1\alpha=1, 0.30.3, 0.20.2, 0.10.1. (Bottom) g=50g=50, α=1\alpha=1, 0.30.3, 0.20.2, 0.10.1. Left-most column corresponds to the isotropic case.

In Figure 12 we provide a more detailed comparison of the cantilever beam designs for the isotropic case α=1\alpha=1 and the strongly anisotropic cases α=0.3,0.2,0.1\alpha{=0.3,0.2,0.1} with loading magnitudes g=50g=50 and g=30g=30. In both situations, the anisotropic designs exhibit less interior structures and connecting bridges with steeper slopes than the isotropic designs, which is to be expected from the construction of γα\gamma_{\alpha}. It is worth mentioning that some (but not all) of the interfacial regions (colored deep blue) are thicker for smaller values of α\alpha. This is attributed to the fact that, from the expression of Φ0\Phi_{0} obtained in the formally matched asymptotic analysis in Section 5.2, the interfacial thickness at a spatial point 𝒙=(𝒔,z)\bm{x}=(\bm{s},z) in the transformed coordinate system is proportional to γ(𝝂(𝒔))\gamma(\bm{\nu}(\bm{s})) with unit normal 𝝂\bm{\nu}. Recalling the left of Figure 3, the normal vectors with directions associated to the line segment LL attain higher values of γ\gamma, and thus a thicker interfacial region, compared to directions associated with the circular arc CC. Smaller values of α\alpha amplify the thickness as the line segment LL is closer to the origin, leading to higher values of γ\gamma.

In Figure 13 we display refined designs taking as initial conditions the final states in Figure 12 as well as a smaller value of ε=1/(64π)\varepsilon=1/(64\pi). The overall designs are similar to Figure 12 with some minor changes, particularly for the isotropic case α=1\alpha=1 with g=50g=50. We also note that the thicknesses of the interfacial layers in Figure 13 are smaller than those in Figure 12, particularly for α=0.1\alpha=0.1, which is due to the smaller values of ε\varepsilon used.

6.3.4 Bridge construction

Our last numerical simulation is inspired by [89, Figure 4], where we use the half-domain setup

Ω\displaystyle\Omega =(0,1)×(12,12),ΓD1={0}×(12,12),ΓD=(78,1)×{12},\displaystyle=(0,1)\times(-\tfrac{1}{2},\tfrac{1}{2}),\quad\Gamma_{D_{1}}=\{0\}\times(-\tfrac{1}{2},\tfrac{1}{2}),\quad\Gamma_{D}=(\tfrac{7}{8},1)\times\{-\tfrac{1}{2}\},
Γg\displaystyle\Gamma_{g} =((0,0.02)(120.02,12+0.02))×{12},𝒈(x1,x2)={(03000)x10.02,(01500)x1>0.02.\displaystyle=((0,0.02)\cup(\tfrac{1}{2}-0.02,\tfrac{1}{2}+0.02))\times\{-\tfrac{1}{2}\},\quad\bm{g}(x_{1},x_{2})=\begin{cases}-\tbinom{0}{3000}&x_{1}\leq 0.02\,,\\ -\tbinom{0}{1500}&x_{1}>0.02\,.\end{cases}

For an elastic material with E=1200E=1200 and ν=0.3\nu=0.3, we investigate the optimal designs with the convex regularized anisotropy (6.2). In Figure 14 we display results for α=1\alpha=1, 0.70.7, 0.50.5 starting from random initial data with mass m=0.04m=0.04. Although the overall shapes are rather similar in all three cases, we observe the anisotropic cases (middle and right figures) have sharper corners on the underside of the bridge, whereas the underside of the isotropic case (left figure) is smoother.

Refer to caption
Refer to caption
Refer to caption
Figure 14: Bridge construction half-domain computation for ε=1/(32π)\varepsilon=1/(32\pi). Anisotropy (6.2) used with δ=0.1\delta=0.1 and α=1\alpha=1 (isotropic case), 0.70.7, 0.50.5. The solutions at time t=0.1t=0.1 starting from random initial data with mass m=0.4m=0.4.

Acknowledgments

The work of KFL is supported by the Research Grants Council of the Hong Kong Special Administrative Region, China [Project No.: HKBU 14302218].

References

  • [1] O. Abdulhammed, A. Al-Ahmari, W. Ameen and S.H. Mian. Additive manufacturing: Challenges, trends, and applications. Adv. Mech. Engrg. 11 (2019), 1–27
  • [2] G. Allaire, M. Bihr and B. Bogosel. Support optimization in additive manufacturing for geometric and thermo-mechanical constraints. Struct. Multidiscip. Optim. 61 (2020) 2377–2399
  • [3] G. Allaire and B. Bogosel. Optimizing supports for additive manufacturing. Struct. Multidiscip. Optim. 58 (2018) 2493–2515
  • [4] G. Allaire, C. Dapogny, R. Estevez, A. Faure and G. Michailidis. Structural optimization under overhang constraints imposed by additive manufacturing technologies. J. Comput. Phys. 351 (2017) 295–328
  • [5] G. Allaire, F. Jouve and A.-M. Toader. Structural optimization using sensitivity analysis and a level-set method. J. Comput. Phys. 194 (2004) 363–393
  • [6] S. Almi and U. Stefanelli. Topology optimization for incremental elastoplasticity: A phase-field approach. SIAM J. Control Optim. 59 (2021) 339–364
  • [7] H.W. Alt. Linear Functional Analysis: An Application-Oriented Introduction. Springer-Verlag London, 2016
  • [8] L. Ambrosio, N. Fusco and D. Pallara. Functions of Bounded Variation and Free Discontinuity Problems. Oxford mathematical monographs. Clarendon Press, 2000
  • [9] 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
  • [10] O. Amir and Y. Mass. Topology optimization for staged construction. Struct. Multidiscip. Optim. 57 (2017) 1679–1694
  • [11] 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
  • [12] V. Barbu. Nonlinear Differential Equations of Monotone Types in Banach Spaces. Springer Science & Business Media, 2010
  • [13] V. Barbu and T. Precupanu. Convexity and Optimization in Banach Spaces. Springer Netherlands, 2012; 4th edition
  • [14] J.W. Barrett, H. Garcke and R. Nürnberg. Numerical approximation of anisotropic geometric evolution equations in the plane. IMA J. Numer. Anal. 28 (2008) 292–330
  • [15] J.W. Barrett, H. Garcke and R. Nürnberg. On the stable discretization of strongly anisotropic phase field models with applications to crystal growth. ZAMM Z. Angew. Math. Mech. 93 (2013) 719–732
  • [16] J.W. Barrett, H. Garcke and R. Nürnberg. Stable phase field approximations of anisotropic solidification. IMA J. Numer. Anal. 34 (2014) 1289–1327
  • [17] J.W. Barrett, H. Garcke and R. Nürnberg. Parametric finite element approximations of curvature driven interface evolutions. Handb. Numer. Anal. 21 (2020) 275–423
  • [18] 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
  • [19] A.C. Barroso and I. Fonseca. Anisotropic singular perturbations - the vectorial case. Proc. Roy. Soc. Edinburgh Sect. A 124 (1994) 527–571
  • [20] G. Bellettini, A. Braides and G. Riey. Variational approximation of anisotropic functionals on partitions. Annali di Matematica 184 (2005) 75–93
  • [21] E. Beretta, L. Ratti and M. Verani. Detection of conductivity inclusions in a semilinear elliptic problem arising from cardiac electrophysiology. Commun. Math. Sci. 16 (2018) 1975–2002
  • [22] 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
  • [23] 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
  • [24] 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
  • [25] L. Blank, H. Garcke, L. Sarbu and V. Styles. Primal-dual active set methods for Allen–Cahn variational inequalities with non-local constraints. Numer. Methods Partial Differential Equations 29 (2013) 999–1030
  • [26] 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
  • [27] G. Bouchitté. Singular Perturbations of Variational Problems Arising from a Two-Phase Transition Model. Appl. Math. Optim. 21 (1990) 289–314
  • [28] B. Bourdin and A. Chambolle. Design-dependent loads in topology optimization. ESAIM: Control Optim. Cal. Var. 9 (2003) 19–48
  • [29] M. Burger and R. Stainko. Phase-field relaxation of topology optimization with local stress constraints. SIAM J. Control Optim. 45 (2006) 1447–1466
  • [30] 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
  • [31] J. Cahn, C.M. Elliott and A. Novick-Cohen. The Cahn–Hilliard equation with a concentration dependent mobility: motion by minus the Laplacian of the mean curvature. Euro. J. Appl. Math. 7 (1996) 287–301
  • [32] 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
  • [33] T.A. Davis, Algorithm 849: a concise sparse Cholesky factorization package. ACM Trans. Math. Software 31 (2005) 587–591
  • [34] C. Dapogny, A. Faure, G. Michailidis, G. Allaire, A. Couvelas and R. Estevez. Geometric constraints for shape and topology optimization in architectural design. Comput. Mech. 59 (2017) 933–965
  • [35] L.C. Evans. Partial Differential Equations. Vol. 19 of Graduate Studies in Mathematics, AMS, 2010
  • [36] L.C. Evans and R. Gariepy. Measure Theory and Fine Properties of Functions. Math. Chem. Ser. CRC Press, Boca Raton, FL. 1992
  • [37] 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
  • [38] F.C. Frank. The geometrical thermodynamics of surfaces. Am. Soc. Metals (1963) 1–15
  • [39] I. Fonseca. The Wulff Theorem revisited. Proc. Roy. Soc. London 432 (1991) 125–145
  • [40] I. Fonseca and S. Müller. A uniqueness proof for the Wulff Theorem. Proc. Roy. Soc. Edinburgh: Sec. A Math. 119 (1991) 125–136
  • [41] H. Garcke, The Γ\Gamma-limit of the Ginzburg–Landau energy in an elastic medium. AMSA 18 (2008) 345–379
  • [42] 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
  • [43] H. Garcke and C. Hecht. Shape and topology optimization in Stokes flow with a phase field approach. Appl. Math. Optim. 73 (2016) 23–70
  • [44] 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
  • [45] 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
  • [46] 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
  • [47] H. Garcke and B. Stinner. Second order phase field asymptotics for multi-component systems. Interfaces Free Bound. 8 (2006) 131–157
  • [48] N. Gardan and A. Schneider. Topological optimization of internal patterns and support in additive manufacturing. J. Manuf. Syst. 37 (2014) 417–425
  • [49] A. Garaigordobil, R. Ansola, J. Santamaría and I. Fernández de Bustos. A new overhang constraint for topology optimization of self-supporting structures in additive manufacturing. Struct. Multidiscip. Optim. 58 (2018) 2003–2017
  • [50] A.T. Gaynor and J.K. Guest. Topology optimization considering overhang constraints: Eliminating sacrificial support material in additive manufacturing through design. Struct. Multidiscip. Optim. 54 (2016) 1157–1172
  • [51] X. Guo, J. Zhou, W. Zhang, Z. Du, C. Liu and Y. Liu. Self-supporting structure design in additive manufacturing through explicit topology optimization. Comput. Methods Appl. Mech. Engrg. 323 (2017) 27–63
  • [52] M.E. Gurtin. Thermomechanics of evolving phase boundaries in the plane. Oxford University Press, 1993
  • [53] A. Hussein, L. Hao, C. Yan, R. Everson and P. Young. Advanced lattice support structures for metal additive manufacturing. J. Mater. Process. Technol. 213 (2013) 1019–1026
  • [54] J. Jiang, X. Xu and J. Stringer. Support Structures for Additive Manufacturing: A Review. J. Manuf. Mater. Process. 2 (2008) 64 (23 pages)
  • [55] R. Kornhuber. Monotone multigrid methods for elliptic variational inequalities II. Numer. Math. 72 (1996) 481–499
  • [56] Y.-H. Kuo, C.-C. Cheng, Y.-S. Lin and C.-H. San. Support structure design in additive manufacturing based on topology optimization. Struct. Multidiscip. Optim. 57 (2018) 183–195
  • [57] K.F. Lam and I. Yousept. Consistency of a phase field regularisation for an inverse problem governed by a quasilinear Maxwell system. Inverse Problems 36 (2020) 045011
  • [58] M. Langelaar. Topology optimization of 3D self-supporting structures for additive manufacturing. Addit. Manuf. 12 (2016) 60–70
  • [59] M. Langelaar. An additive manufacturing filter for topology optimization of print-ready designs. Struct. Multidiscip. Optim. 55 (2017) 871–883
  • [60] M. Langelaar. Combined optimization of part topology, support structure layout and build orientation for additive manufacturing. Struct. Multidiscip. Optim. 57 (2018) 1985–2004
  • [61] M. Leary, M. Babaee, M. Brandt and A. Subic. Feasible build orientations for self-supporting fused deposition manufacture: a novel approach to spacefilling tessellated geometries. Adv. Mater. Res. 633 (2013) 148–168
  • [62] M. Leary, L. Merli, F. Torti, M. Mazur and M. Bandt. Optimal topology for additive manufacture: A method for enabling additive manufacture of support-free optimal structures. Mater. Des. 63 (2014) 678–690
  • [63] J. Liu and H. Yu. Self-Support Topology Optimization With Horizontal Overhangs for Additive Manufacturing. J. Manuf. Sci. Eng. 142 (2020) 091003 (14 pages)
  • [64] L. Lu, A. Sharf, H.S. Zhao, Y. Wei, Q.N. Fan, X.L. Chen, Y. Savoye, C. Tu, D. Cohen-Or and B. Chen. Build-to-Last: Strength to Weight 3D Printed Objects. ACM Trans. Graph. 33 (2014) 1–10
  • [65] Y. Luo, O. Sigmund, Q. Li and S. Liu. Additive manufacturing oriented topology optimization of structures with self-supported enclosed voids. Comput. Methods Appl. Mech. Engrg. 372 (2020) 113385
  • [66] F. Mezzadri and X. Qian. A second-order measure of boundary oscillations for overhang control in topology optimization. J. Comput. Phys. 410 (2020) 109365
  • [67] A.M. Mirzendehdel and K. Suresh. Support Structure Constrained Topology Optimization for Additive Manufacturing. Comput. Aided Des. 1 (2016) 1–13
  • [68] L. Modica. The gradient theory of phase transitions and minimal interface criterion. Arch. Rat. Mech. Anal. 98 (1987) 123–142
  • [69] L. Modica and S. Mortola. Un esempio di Γ\Gamma-convergenza. (Italian) Boll. Un. Mat. Ital. B (5) 14 (1977) 285–299
  • [70] H.D. Morgan, J.A. Cherry, S. Jonnalagadda, D. Ewing and J. Sienz. Part orientation optimisation for the additive layer manufacture of metal components. Int. J. Adv. Manuf. Technol. 86 (2016) 1679–1687
  • [71] N.C. Owen and P. Sternberg. Nonconvex variational problems with anisotropic perturbations. Nonlinear Anal. 16 (1991) 705–719
  • [72] J. Pellens, G. Lombaert, B. Lazarov and M. Schevenels. Combined length scale and overhang angle control in minimum compliance topology optimization for additive manufacturing. Struct. Multidiscip. Optim. 59 (2019) 2005–2022
  • [73] 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
  • [74] H. Proff and A. Staffen. Challenges of Additive Manufacturing. Why companies don’t use Additive Manufacturing in serial production. Deloitte report (2019), https://www2.deloitte.com/content/dam/Deloitte/de/Documents/operations/Deloitte_Challenges_of_Additive_Manufacturing.pdf
  • [75] X. Qian. Undercut and Overhang Angle Control in Topology Optimization: a Density Gradient based Integral Approach. Int. J. Numer. Methods Eng. 111 (2017) 247–272
  • [76] C. Rupprecht. Projection type methods in Banach space with application in topology optimization. PhD thesis. University of Regensburg, Regensburg, 2016
  • [77] 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
  • [78] 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
  • [79] J.E. Taylor. Existence and structure of solutions to a class of nonelliptic variational problems. Sympos. Math. 14 (1974) 499–508
  • [80] J.E. Taylor. Unique structure of solutions to a class of nonelliptic variational problems. Proc. Sympos. Pure Math. 27 (1975) 419–427
  • [81] J.E. Taylor and J.W. Cahn. Linking anisotropic sharp and diffuse surface motion laws via gradient flows. J. Statist. Phys. 77 (1994) 183–197
  • [82] D. Thomas. The development of design rules for selective laser melting. University of Wales Institute, PhD thesis (2009)
  • [83] C.-J. Thore, H. Alm Grundström, B. Torstenfelt and A. Klarbring. Penalty regulation of overhang in topology optimization for additive manufacturing. Struct. Multidiscip. Optim. 60 (2019) 59–67
  • [84] F. Tröltzsch. Optimal Control of Partial Differential Equations. Theory, Methods and Applications. Grad. Stud. in Math. 112, AMS, Providence, RI, 2010
  • [85] E. van de Ven, R. Maas, C. Ayas, M. Langelaar and F. van Keulen. Continuous front propagation-based overhang control for topology optimization with additive manufacturing. Struct. Multidiscip. Optim. 57 (2018) 2075–2091
  • [86] E. van de Ven, R. Maas, C. Ayas, M. Langelaar and F. van Keulen. Overhang control based on front propagation in 3D topology optimization for additive manufacturing. Comput. Methods Appl. Mech. Engrg. 369 (2020) 113169
  • [87] J. Vanek, J.A.G. Galicia and B. Benes. Clever support: Efficient support structure generation for digital fabrication. Comput. Graph. Forum 33 (2014) 117–125
  • [88] C. Wang and X. Qian. Simultaneous optimization of build orientation and topology for additive manufacturing. Addit. Manuf. 34 (2020) 101246
  • [89] M.Y. Wang and S. Zhou. Phase field: a variational method for structural topology optimization. CMES Comput. Model. Eng. Sci. 6 (2004) 547–566
  • [90] G. Wulff. Zur Frage der Geschwindigkeit des Wachstums und der Auflösung der Kristallflächen. Z. Krist. 34 (1901) 449–530
  • [91] X. Zhang, X. Le, A. Panotopoulou, E. Whiting and C.C.L. Wang. Perceptual models of preference in 3D printing direction. ACM Trans. Graph. 34 (2015) 1–12
  • [92] D. Zhao, M. Li and Y. Liu. A novel application framework for self-supporting topology optimization. Vis. Comput. 37 (2021) 1169–1184