476
Views
11
CrossRef citations to date
0
Altmetric
Articles

Shape optimization of a breakwater

ORCID Icon & ORCID Icon
Pages 936-956 | Received 11 Dec 2014, Accepted 25 Jul 2015, Published online: 23 Sep 2015

Abstract

In this paper, we optimize the shape of a breakwater which protects a harbour basin from incoming waves. More specifically, our objective is reducing the harbour resonance due to long-range ocean waves. We consider the complex-valued Helmholtz equation as our model state equation and minimize the average wave height in the harbour basin with the shape of the breakwater as optimization variable. The geometry is described by the level set method, i.e. the domain is given as the subzero level set of a function. In contrast to many publications we use the volume expression of the shape derivative, which lends itself naturally to a level set update via a transport equation. The model problem features intrinsic geometric constraints which we treat in the form of forbidden regions. We guarantee feasibility of the iterates by projecting the gradient onto a suitable admissible set.

AMS Subject Classifications:

1. Introduction

Since the introduction of the level set method to shape optimization at the turn of the century it has developed into one of the most powerful techniques in this area. Some early works are [Citation1Citation5]. Many publications deal with structural optimization, usually the Ersatz material approach in the hold-all domain D is used to compute the mechanical properties of the structure and the domain Ω is never explicitly resolved. But there are also approaches where the domain is exactly meshed, consult for instance.[Citation6Citation12] The literature is extensive, we refer to the review paper [Citation13] for an overview of level set-based methods in structural topology and shape optimization. The survey [Citation14] presents the level set method in combination with inverse problems and optimal design. For a more comprehensive exposition of the broad field of shape optimization we mention the monographs.[Citation15Citation23]

In this paper, our objective is to reduce the resonance of a harbour due to long range ocean waves. In practice many publications (e.g. [Citation24Citation28]) employ the mild slope equation to model wave effects in coastal areas. It was first derived by Berkhoff [Citation29] and can be written as(1) ·CCgφ+k2CCgφ=0,(1)

where φ is the horizontal variation in velocity potential, k is the wave number, ω is the wave frequency, C=ω/k is the wave celerity, Cg=C21+2khsinh2kh is the group velocity and h is the water depth. Usually it is enriched with additional effects such as partial absorption boundaries, bottom friction, entrance loss, etc. However, the aim of this paper is to introduce the necessary steps for solving the associated shape optimization problem with the level set method and not a realistic modelling of harbour resonance. For this reason, we assume as simplification that the water depth is constant throughout, which leads to the well-known Helmholtz equation(2) Δφ+k2φ=0.(2)

Note that the potential φ in this case is complex valued. To the best of our knowledge such a shape optimization problem was only briefly considered in the thesis [Citation30]. The author used the real-valued Helmholtz equation as state equation, an explicit discrete geometry description via the finite element mesh and computed the shape derivative by differentiating with respect to nodal coordinates.

We will consider the complex-valued Helmholtz equation as our model state equation and want to minimize the average wave height in the harbour basin with the shape of the breakwater as optimization variable. The geometry will be described by the level set method,[Citation31Citation34] i.e. the domain Ω is given as the subzero level set of a function Φ:DR, where DR2 is the hold-all domain. In the numerical realization the level set function is given on a regular grid and we approximate Ω with one or more polygonal curves. In each iteration, we construct a triangulation of D which resolves the interface and use it to compute the state, adjoint state and gradient of the cost functional. The model state equation is given on an exterior domain. Since we want to use finite elements we follow the presentation in [Citation35] and reformulate the problem on a bounded domain. In contrast to many publications, we use the volume expression of the shape derivative, which lends itself naturally to a level set update via a transport equation. It requires less regular finite element functions and in our experience the volume expression is numerically more stable than the Hadamard form. This assessment is shared in the recent papers [Citation36,Citation37]. In [Citation37], the volume expression of the shape derivative and the level set method are also used.

The model problem naturally involves geometric constraints in the form of forbidden regions. We strictly enforce those constraints by projecting the gradient onto a suitable admissible set. From a theoretic point of view the scalar product used for the projection should be at least as smooth as the scalar product used to determine the gradient. In the numerical experiments we will study different choices of gradient and projection and see that sometimes a less regular projection leads to better results.

2. Description of the physical model

We study the situation depicted in Figure . We have an isle bounded by the contour ΓL, some breakwaters given by ΓB and a surrounding ocean denoted by Ω+. We want to compute the scattered wave u induced by an incoming planar monochromatic wave z(x)=exp(ikwTx) with incident direction wR2 and wave number k>0. The total surface perturbation is given by y=u+z. We study the model(3) Δy+k2y=0inΩ+ay+ny=0onΓI:=ΓBΓL.(3)

Here a:R2C describes the absorption coefficient at the boundary. The boundaries ΓL,ΓB are assumed to be Lipschitz. Let us discuss appropriate boundary conditions. It is a standard assumption that the scattered wave satisfies the Sommerfeld radiation conditioniku-Ru=o(R-12),forR.

Figure 1. The domain Ω.

Figure 1. The domain Ω.

If we want to study this problem in a weak form on the unbounded domain Ω+ we would need to introduce different weighted Sobolev spaces for the test and ansatz functions and include the Sommerfeld radiation condition in the ansatz space. See [Citation38, Chapter 4] and [Citation35, Section 2.3] for more details. This approach leads, however, to various difficulties in the numerical realization if a finite element discretization is employed. While it would be possible to handle the Helmholtz equation on an exterior domain with the boundary element method, an extension to the more realistic mild slope equation would be very challenging. For this reason we focus on the finite element approach.

An alternative is to decompose the domain Ω+ disjointly into a bounded domain Ω and an unbounded domain Ωa by introducing an artificial smooth boundary Γa, such that Ω+=ΩΓaΩa. The problem (Equation3) is then equivalent to the following coupled problem (cf. [Citation39])(4) Δy-+k2y-=0inΩay-+ny-=0onΓIy-=y+onΓany-=ny+onΓaΔy++k2y+=0inΩaiku+-Ru+=o(R-12)forR.(4)

For a given u- on Γa one can solve the unbounded Dirichlet problem (recall u=y-z)Δu++k2u+=0inΩau+=u-onΓaiku+-Ru+=o(R-12)forR,

compare [Citation38]. If we have the solution u+ we can easily compute nu+=nu-onΓa. We denote the mapping u-nu- by Ge and observe that GeL(H12(Γa),H-12(Γa)). This operator is called the Dirichlet-to-Neumann operator. There exists an integral and a series representation of the non-local operator Ge, the integral variant is also called the Steklov–Poincaré operator. The Neumann condition ny-=ny+onΓa can be reformulated byny-=nu-+nz=Geu-+nz=Gey--Gez+nz.

We replace Ge by some yet unspecified GL(H12(Γa),H-12(Γa)), which might be some approximation of the exact solution operator Ge. Hence we arrive at the bounded problem(5) Δy+k2y=0inΩay+ny=0onΓIy=Gy-Gz+nzonΓa,(5)

which is equivalent to (Equation3) for the choice G=Ge.

Let us fix some conventions and notation at this point. We define the usual bilinear L2-scalar product for real-valued functions on some set BRn as (·,·)L2(B) and the corresponding sesquilinear form as (f,g)LC2(B):=(f,g¯)L2(B) for some complex valued functions f,g. Furthermore we introduce the real-valued scalar product(f,g)LR2(B):=Re(f,g)LC2(B).

The norm ||·||LR2(B) induced by this scalar product coincides with the norm induced by the sesquilinear form. Hence the elements of the spaceLR2(B):={f:AC|||f||LR2(B)<}

coincide with the elements of {f:BC|||f||LC2(B)<}, but since we use the (·,·)LR2(B) scalar product we have a different Hilbert space structure. Other Hilbert spaces will be treated analogously (e.g. HR1(B)).

Let us get back to the model problem. We make the following simplifying assumption:

Assumption 2.1

We choose G as the 0th-order approximation of GeGy,φHR-12(Γa),HR12(Γa)=(iky,φ)LR2(Γa),

cf. [Citation35, Section 3.]. Furthermore the absorption coefficient is set to a0 (i.e. perfect reflection).

Remark 2.2

(1)

The choice of G is a rather crude simplification, but in this paper our focus is more on methodology and not so much on realistic modelling. The low-order approximation of Ge introduces artificial reflections at Γa, but since this makes the shape optimization problem presumably harder to solve we accept that for the moment. For more evolved methods of treating the artificial boundary Γa and the operator Ge we refer to [Citation35, Chapter 3].

(2)

Setting the absorption coefficient to zero implies that there is no damping effect by absorption of energy at the reflecting boundary. We note again that this presumably makes the optimization problem harder to solve since small design changes might have large non-local effects due to wave interference. Furthermore, while breakwaters are efficient in reducing the wave amplitude for waves with short wave periods (like wind waves), this does not hold for long wave periods. Harbour oscillations and resonance due to long waves has been widely studied in coastal engineering literature, see for example [Citation40] and the references therein.

The weak formulation of (Equation5), with G and a chosen to satisfy Assumption 2.1, reads(6) FindyHR1(Ω):b(y,φ)=f(φ),φHR1(Ω),(6)

where we define(7) b(y,φ):=(y,φ)LR2(Ω)-k2(y,φ)LR2(Ω)-(iky,φ)LR2(Γa),f(φ):=(nz-ikz,φ)LR2(Γa).(7)

Results concerning the existence of a unique solution of (Equation6) and its regularity are well known:

Theorem 2.3

Let Ω be a Lipschitz domain. Then there exists a unique solution yHR1(Ω) of (Equation6) for any right-hand side fHR1(Ω) and we have ||y||HR1(Ω)c||f||HR1(Ω) for some c>0.

Proof

We have the Gelfand-triple HR1(Ω)LR2(Ω)HR1(Ω). Further b(·,·) is HR1(Ω)-coercive. Hence the Fredholm alternative holds: Either there exists a unique solution of (Equation6) for any fHR1(Ω) or there exists a non-trivial solution y00 of the homogenous problemb(y,φ)=0,φHR1(Ω).

For our choice of G the solution of the homogenous problem is unique [Citation35, Theorem 3.2].

Theorem 2.4

Let Ω be a C2-domain. The solution yHR1(Ω) of (Equation6) has the additional regularity yHR2(Ω).

Proof

Follows from standard regularity results for elliptic equations cf. [Citation41, Theorem 9.1.20].

3. Shape optimization problem

As announced in the introduction our objective is to minimize the average wave height in the harbour basin Q (compare Figure ). Given a solution y of (Equation6) we define the cost functional asJ(y)=12||y||LR2(Q)2.

Obviously enclosing the whole harbour basin by a breakwater is not a feasible solution, so we have to introduce an harbour approach A and demand that QA is always part of the ocean. Furthermore, we do not want to remove parts of the island (the inhabitants might complain). As usual in shape optimization we restrict our considerations to a holdall domain DR2 which contains all relevant shapes. In our setting it is natural to choose D such that D=Γa. Hence, we define the admissible set of domains to be(8) Oad:={ΩD|(QA)Ω,ΩL=},(8)

where ΩOad represents the ocean. Of course one could impose additional constraints, e.g. a volume constraint on the set of admissible domains. We can now formulate the abstract shape optimization problemminimizeJ(y)such thatΩOad,andysolves(6) onΩ.

In this paper, we will not concern ourselves with the question of existence of a solution of the shape optimization problem. For a detailed discussion of techniques and conditions which guarantee the existence and uniqueness of solutions to shape optimization problems, we refer to [Citation18,Citation19]. In order to find a solution (if it exists) we develop a framework for a gradient descent method. For this we follow the usual optimal control approach of introducing a design-to-state mapping Ωy(Ω) and computing the derivative of the reduced objective j(Ω):=J(y(Ω)). In the next section, we outline the theory of computing shape derivatives and apply it to our problem setting.

4. Shape sensitivity analysis

Following the presentation in [Citation18,Citation42] we introduce a bounded hold-all Lipschitz domain DRn, D, which contains all relevant geometrical objects, and study vector fields V:D¯Rd, satisfying the conditions(9) Vis globally Lipschitz continuous inD¯,and supp(V)D.(9)

Note that in [Citation18,Citation42] a more general setting is considered, but this suffices for our purposes. If the velocity field V satisfies condition (Equation9) we associate it with the flow mapTV:[0,τ]×D¯Rnwhich satisfiestTV(t,x)=V(x).

We abbreviate TV(t):=TV(t,·). The flow map transforms a set ΩD¯ into a new domainTV(t)(Ω)={TV(t,x)|xΩ},

which is also contained in D¯. Furthermore, we introduce the spaces of k-times differentiable functions RnRn with compact support in D, i.e.Cck(D,Rn):={VCk(D,Rn)|suppVcD}.

If V satisfies (Equation9) the transformation TV(t) is bi-Lipschitz for all τ>0 and t[0,τ], cf. [Citation18, Theorem 4.2.1]. If VCck(D,Rn) then TV(t) is a Ck-diffeomorphism.

Recall the classical notions of the Eulerian semiderivative and shape differentiability:

Definition 4.1

[Citation42, Definitions 3.1 and 3.2] Let OP(D) be an appropriate set of shape variables contained in the hold-all domain DRn.

(1)

The Eulerian semiderivative of a shape functional j:OR at ΩO in a direction V satisfying (Equation9) is given bydj(Ω;V):=limt0j(TV(t)(Ω))-j(Ω)t, if the limit exists and is finite.

(2)

The shape functional j:OR is said to be shape differentiable at ΩO, if the Eulerian semiderivative exists for all VCc(D,Rn) and the mapdj(Ω;·):Cc(D,Rn)R,Vdj(Ω;V), is linear and continuous.

Remark 4.2

(1)

The Hadamard-Zolesio structure theorem (cf. [Citation42, Theorem 3.2]) shows that the support of the vector distribution dj(Ω;·)Cc(D,Rn) is contained in ΩD.

(2)

The proofs of most formulas for the computation of shape derivatives rely on the representation of the shape functional j in terms of the transformation TV(t). If j depends on the integral over a n-dimensional subset of TV(t)(Ω) and a shape-depended state y the directional derivative has a natural representation as a volume integral. If the boundary Ω is smooth enough this can be related via the Gauß divergence theorem to a boundary representation in accordance with the structure predicted by the Hadamard-Zolesio theorem.

(3)

If the boundary Ω of ΩintD is compact then dj(Ω;·) is continuous for the Cck(D,Rn)-topology for some k0, and dj(Ω;·)H-s(D,Rn) for some s0 (cf. [Citation18, Remark 9.3.1]).

(4)

The derivative dj(Ω;·)Cc(D,Rn) is often called the shape gradient. We think that the terms derivative and gradient should be clearly separated. Whereas, the derivative is an element of the dual space of some vector space, in our terminology the gradient is the Riesz representative of the derivative with respect to some scalar product. So if dj(Ω;·)H for some Hilbert space H the gradient j(Ω) with respect to the H-scalar-product is given by(j(Ω),V)H=dj(Ω;V),VH.

Let us demonstrate how the shape derivative of j(Ω)=J(y(Ω)) in a direction V satisfying (Equation9) can be calculated. This is not trivial since the states y(TV(t)(Ω))HR1(TV(t)(Ω)) depend on the varying domains and are defined in function spaces which also depend on t. The standard approach to this problem is to introduce a function space parameterization. It is based on the equivalencegHR1(TV(t)(Ω))gTV(t)HR1(Ω)

which goes back to [Citation43, Section 2.3.1]. Hence we can transform the state equation given on TV(t)(Ω) onto the current domain Ω. This operation is also called pull back.

We introduce the notationDTV(t):=Jacobian matrix ofTV(t),A(t):=det(DTV(t))DTV(t)-1DTV(t)-T,

and note that det(DTV(t))=|det(DTV(t))|fort0 small. Furthermore V=0 on Γa by (Equation9) and hence TV(t)(Γa)=Γa for all t0. Consider now the solution yt of the state Equation (Equation6) on the varying domains TV(t)(Ω) for t[0,τ]. We transform the left-hand sidebt(yt,φt):=(yt,φt)LR2(TV(t)(Ω))-k2(yt,φt)LR2(TV(t)(Ω))-(ikyt,φt)LR2(TV(t)(Γa))

onto the reference domain via yt=ytTV(t), φt=φtTV(t), andb(t,yt,φt):=(A(t)yt,φt)LR2(Ω)-k2(det(DTV(t))yt,φt)LR2(Ω)-(ikyt,φt)LR2(Γa).

The boundary conditions are defined on the fixed boundary Γa and do not change through the transformationf(t,φt)=zn-ikz,φtLR2(Γa)=zn-ikz,φ0LR2(Γa)=f(φ0).

Hence, the state yt:=ytTV(t)HR1(Ω) solves the transformed state equation(10) b(t,yt,φ)=f(t,φ),φHR1(Ω).(10)

We also need to transform the objective. Due to (Equation8) we consider only velocity fields satisfying V=0 on Q. Introducing j~:[0,τ]R with j~(t):=12||yt||LR2(Q)2 it holdsj~(t)=12||yt||LR2(Q)2=J(yt)=j(TV(t)(Ω)).

Due to [Citation18, Theorem 9.3.4] the shape derivative of j at Ω can be obtained asdj(Ω;V)=j~(0).

Remark 4.3

Calculating the derivative of j~ amounts to calculating the material derivative y˙t:=limt0(yt-y0)/t. An alternative approach would be the local shape derivative of yt, but in our opinion the material derivative approach is more general and canonical. As usual in PDE-constrained optimization if one is interested in the whole derivative j(Ω) the direct approach of calculating directional derivatives is infeasible. Instead it is advantageous to employ the adjoint approach. We demonstrate this now briefly.

Introducing the Lagrange functionalL:[0,τ]×HR1(Ω)×HR1(Ω)R:L(t,y,p)=J(y)+b(t,y,p)-f(p),

we note that for a solution ytHR1(Ω) of the state equation it holdsj~(t)=L(t,yt,p),pHR1(Ω).

Furthermore, if ptHR1(Ω) solves the adjoint equation(11) yL(t,yt,pt),ψHR1(Ω),HR1(Ω)=(yt,ψ)LR2(Q)+b(t,ψ,pt)=0,ψHR1(Ω),(11)

thendj(Ω;V)=j~(0)=tL(0,y0,p0)=tb(0,y0,p0)

Note that tdet(DTV(0))=div(V) andtA(0)=div(V)I-DV-DVT,

where I denotes the identity matrix in R2, cf. [Citation18, Section 9.4.1]. Hence we obtain the following volume representation of the shape derivative(12) dj(Ω;V)=(A(0)y0,p0)LR2(Ω)-k2(div(V)y0,p0)LR2(Ω).(12)

It is easy to check that dj(Ω;·)W1,(D,R2) and hence dj(Ω;·)Hs(D,R2) for s>2. Note that we can choose a bounded, open Lipschitz holdall domain DR2. In this case, W1,(D,R2)=C0,1(D,R2) algebraically and topologically.

5. Geometric constraints

In this section we describe how to incorporate geometric constraints in the shape optimization problem. Geometric constraints appear frequently in applications. In our breakwater optimization example we have the following: the harbour basin Q and the harbour approach A should always be water regions, i.e. be a part of Ω. On the other hand the island L should never be a part of Ω.

A very natural idea for solving the constrained optimization problem(13) minimizej(Ω)such thatΩOad,(13)

is to use appropriate projections. The projected descent method is the prototypical example of an algorithm which always stays feasible with regard to the constraints. The idea is to take a step in an appropriately chosen descent direction and afterwards to project onto the admissible set. The method is classical in the context of optimization in Hilbert spaces (cf. [Citation44, Section 2.2.2] and has recently also been generalized to the Banach space context (cf. [Citation45]). Unfortunately an extension to the shape optimization framework is not straightforward. While one can define a metric on the group of images of a domain ([Citation18, Chapter 3]) the practical realization of such the associated projection is a challenging topic. We leave this option open and note that it would be a very interesting topic for further research. There are also other approaches to define metrics on families of sets, some of them even feature associated projections which are easy to implement. However, for these metrics there is no intrinsic notion of an appropriate general tangent space available. Hence, the computation of derivatives which are compatible with these metrics can only be done in specific situations, but not in a canonical way.

We propose to use an alternative approach. We project the descent direction onto a suitable set of velocity fields which keep the deformed domain in the admissible set. The easiest way to ensure this is to defineVfeas:={VC0,1(D¯,R2)|V|F=0,F=QALD}.

Obviously if ΩOad and VVfeas, then TV(t)(Ω)Oad for all t>0. There is some freedom in defining the descent direction and the associated projection. Note that, in general, these two quantities should be derived with respect to the same scalar product. Otherwise one cannot guarantee that the projected direction is a descent direction. There are already easy examples in finite dimension showing that the projected Newton direction is not necessarily a descent direction. Hence we consider

Assumption 5.1

HC0,1(D¯,R2) is a Hilbert space and (·,·)a:H×HR is a symmetric, continuous, and coercive bilinear form with associated norm ||·||a.

Remark 5.2

(1)

We work here with an imbedded Hilbert space instead of the Banach space C0,1(D,Rd). It might be interesting to see whether one can transfer the results of [Citation45] also to our setting.

(2)

One can easily generalize the fixed bilinear form (·,·)a to a variable one and retain convergence of the corresponding algorithm under suitable uniformness requirements, cf. e.g. [Citation45]. Thus it is possible to consider second-order information in the bilinear form.

Definition 5.3

Let Assumption 5.1 be satisfied and denote Vad:=VfeasH. Define the projection with respect to (·,·)aPa:HVad:Pa(V)Vad,and(V-PaV,W)=0WVad,

and for ΩkOad the directionVkH:(Vk,W)a=dj(Ωk;W)WH.

Remark 5.4

We will in the following refer to Vk as gradient, since it is the Riesz representative of dj(Ωk;·) with respect to (·,·)a. In particular -Vk is the direction of steepest descent w.r.t. ||·||a. Note again, that with a suitable choice of the bilinear form one could also obtain a Newton-type direction, or some other gradient related direction.

Since Vad is a linear subspace the projection can be implemented efficiently. In fact, instead of first computing the gradient Vk and then projecting it we can directly compute the gradient with respect to the subspace and obtain the same velocity field.

Lemma 5.5

Let Assumption 5.1 be satisfied. The element UkVad satisfies(14) (Uk,W)a=dj(Ωk;W)WVad(14)

if and only if there holds Uk=Pa(Vk).

Proof

By definition of the gradient it holds for all WH(Vk-Uk,W)a=(Vk,W)a-(Uk,W)a=dj(Ωk;W)-(Uk,W)a.

Considering WVad shows the equivalence of the statements.

We have the following result concerning the optimality of the current iterate Ωk.

Lemma 5.6

Let Assumption 5.1 be satisfied.

(1)

If Pa(Vk)=0, then it holds dj(Ωk;W)=0 for all WVad.

(2)

If Pa(Vk)0, then Pa(-Vk) is a descent direction, i.e.dj(Ωk;Pa(-Vk))=-||Pa(Vk)||a2<0.

Proof

This follows directly from the definitions, respectively, from the above equivalence.

Remark 5.7

Note that Lemma 5.6 does not guarantee us, that Ωk is a local solution of (Equation13) if Pa(Vk)=0. We can only expect to obtain a local solution of the restricted problemminimizej(Ω)such thatΩ{TV(t)(Ω0)|VVad,t0}.

In that sense we are only computing a suboptimal solution of the original problem.

6. Level set method

The level set method was introduced in [Citation31] and is widely used to describe propagating fronts, moving interfaces, image segmentation, morphing bodies and many similar applications. We refer to the monographs [Citation32,Citation34] for a detailed presentation of this rich topic. The core idea is to describe some ΩRn as the subzero level setΩ:={xRn|Φ(x)<0}

of some continuous function Φ:RnR. Denote Γ={xRn|Φ(x)=0} and assume Γ=Ω. For ΦCk(Rn,R), k1, Γ and assuming Φ(x)0xΓ, it holds that Γ is a Ck-submanifold of dimension (n-1) in Rn.

Let V satisfy (Equation9). A family of domains TV(t)(Ω) can be encoded via a family of level set functions. SettingΦ:[0,τ]×Rn:Φ(t,x)=Φ0(TV(t)-1(x)),

where Φ0 is a level set function of Ω, it holds TV(t)(Ω)={xRn|Φ(t,x)<0}. Indeed it is a classical result that a solution of the transport equation(15) tΨ+ΨTV=0with initial conditionΨ(0,x)=Φ0(x),(15)

where ΨTV is meant in the weak sense, satisfies for a.e. xRn and all t[0,τ]Ψ(t,x)=Φ0(TV(t)-1(x),

cf. [Citation46, Proposition 3.3]. Note that this result is due to the regularity of V. In general the transport equation may lead to shocks, discontinuous solutions and special solution concepts are required. We point out in particular the concept of viscosity solutions, cf. e.g. [Citation33,Citation47].

Summarizing if the velocity field along which we want to transform the current domain Ωk is given by the projected negative gradient Pa(-Vk) we can use (Equation15) to obtain a representation of the next iterate level set function and thus of Ωk+1.

7. Optimization and discretization aspects

Let us briefly sketch Algorithm 7.1 before going into details. Starting with a level-set function Φ given on a regular grid we extract a discretization of the domain Ω and the interface ΓI=ΓBΓL. We solve the state and adjoint equations on the discretized domain using piecewise linear finite elements. Now we can compute the shape derivative of the reduced objective and obtain a projected gradient representation from (Equation14). Finally the level set function is evolved according to (Equation15) along the negative projected gradient for some time span Δt which we choose such that it satisfies the Armijo condition.

We approximate the interface ΓI, given by the zero level-set of Φ, by one or multiple polygonal curves. Between each pair of neighbouring points on the regular grid at which the level set function has different signs there is an intersection point of the zero level set with the edges of the grid. We approximate this intersection point using an affine model for Φ along the edge. Connecting all these intersection points, we obtain the polygonal approximation ΓI,h to ΓI and thus the current domain Ωh.

In the next step, the domains Ωh and D are discretized with triangular meshes which resolve the polygonal boundary ΓI,h. Furthermore, the mesh representing Ωh consists of a subset of the triangles of the mesh for D. Cells of the rectangular grid for which all four vertices have the same sign of Φ are split along their diagonal into two triangles, and cells which are intersected by ΓI,h are split depending on how they are intersected. In order to avoid triangles that are too degenerate and cause numerical difficulties, we enforce a certain minimum ratio between edges of all mesh triangles.

The mesh on Ωh is used to solve the state and adjoint equations, and the mesh on D is used to solve (Equation14) for the projected gradient. Our mesh construction ensures that each point of the original regular grid is also a vertex of the triangle mesh, so that we can extract Pa(-Vk) on each grid point to solve (Equation15).

Let us briefly comment on the discretization of the state and adjoint equation. We use the usual machinery of piecewise linear continuous finite elements to discretize the scalar products (f,g)L2(Ωh),(f,g)L2(Ωh),and(f,g)L2(Γa) for some real valued functions f,g. Recall the notation from Section 2. Splitting every complex valued function f into f=f1+if2, with fi:ΩhR, we find(f,g)LR2(Ωh)=Re(f1+if2,g1+ig2)LC2(Ωh)=(f1,g1)L2(Ωh)+(f2,g2)L2(Ωh).

Analogously it holds that(f,g)LR2(Ωh)=(f1,g1)L2(Ωh)+(f2,g2)L2(Ωh),

and(16) (if,g)LR2(Γa)=(f1,g2)L2(Γa)-(f2,g1)L2(Γa).(16)

Note that z(x)=exp(ikwTx) which implies nz-ikz=ik(wTn-1)z. Hence we only need to implement the boundary expression (Equation16). Combining these formulas with the usual mass and stiffness matrices we can easily assemble the system matrix and right-hand side of the state equation (Equation10) and adjoint equation (Equation11).

Note that, in order to guarantee the minimum regularity of C0,1(D,R2) for the velocity field, we would need to compute the gradient in Hs(D,R2) with s>2. In our numerical experiments we neglect to do so. For ease of implementation we only use piecewise linear continuous finite elements to discretize the ansatz and test spaces of (Equation14). We employ the triangulation of D as described above. We experiment with different choices of the scalar product used to compute the gradient and the projection, cf. Section 8. The derivative of the reduced objective in the direction V is given bydj(0;V)=(A(0)y,p)LR2(Ωh)-k2(div(V)y,p)LR2(Ωh).

Using simple matrix manipulations we can rewrite this expression such that the discretized version djh(0) of the derivative can be evaluated by a simple vector multiplication djh(0)TV.

Once we have computed the projected gradient, we need to update our geometry. For this we solve(17) tΨ(t)+UTΨ(t)=0t(0,Δt),Ψ(0)=Φ,(17)

with U=Pa(-Vk) and use Ψ(Δt) as the new level set function in the next iteration. The time span Δt is determined by the Armijo rule, i.e.(18) j~(Δt)j~(0)+Δtγj~(0).(18)

As was suggested in [Citation37], we employ the Local Lax-Friedrichs flux (cf. [Citation48]) and an explicit Euler time stepping scheme to evolve Ψ. In our setting this leads to the level set function updates(19) Ψijl+1=Ψijl-δtHLLF.(19)

Here Ψijl is the value of the level set function in the node (xi,yj) of the regular grid at the lth time step,HLLF=p-+p+2U1+q-+q+2U2-12(p+-p-)|U1|-12(q+-q-)|U2|,

andp-=Ψijl-Ψi-1,jlΔx,p+=Ψi+1,jl-ΨijlΔx,q-=Ψijl-Ψi,j-1lΔy,q+=Ψi,j+1l-ΨijlΔy.

The time step size δt is chosen to satisfy the Courant-Friedrichs-Lewy condition which guarantees the stability of the time stepping scheme. If many time steps are necessary it is common to reinitialize the level set function every few steps for numerical stability. We use the signed distance function of the current domain (as defined by Ψl) which we obtain via a fast marching algorithm.[Citation49] In our numerical experiments the reinitialization was only rarely necessary.

8. Numerical experiments

We implemented the proposed projected gradient method in GNU Octave.[Citation50] The routines for generating the geometry from the level set function and assembling the finite element mesh are using the Octave package level-set.[Citation51] It also provides a method to compute the signed distance function using a fast marching algorithm.

In our numerical experiments the harbour basin Q corresponds to the box |x|0.4, and -0.3y0.0. The harbour approach is given by |x|0.1,0y and the isle by |x|0.55,-0.6y-0.4. We present the results for two different initial layouts of the breakwater ΓB, cf. Figure . The computational effort is reduced by imposing that everything outside the box -1x,y1 is ocean. We study an incoming wave with wave number k=7 and compare the effects of different scalar products for the computation of the projected gradient. We present results for the choices (·,·)a=(·,·)H1(D) and an approximation of the bi-Laplacian scalar product, i.e.(·,·)a(·,·)L2(D)+(Δ·,Δ·)L2(D).

Furthermore, we experiment with an H1 gradient which is projected with respect to the L2 scalar product. Note that in this case we are not guaranteed to have a descent direction, cf. Section 5. However, this choice also has some advantages and, with a slight modification, performs sometimes better than the other two variants.

For the first experiment the regular grid consists of 501×501 points and the incident wave direction is w=(0.5,-0.5)T. The initial geometry can be seen on the left side of Figure . Note the wave resonance in the harbour basin. In order to be able to compare the results, the L2 norm of the projected gradient is used as performance criterion in all examples. Note that optimization and discretization do not commute in our approach. Furthermore, we are only using a first-order method, hence convergence towards a critical point of the discrete problem cannot be expected in a reasonable number of steps. We stop the algorithm either if the L2 norm of the gradient is below 1.0e-5 or after 3000 iterations. In general, the performance of the different methods reflects the smoothness of the different gradients. Whereas the bi-Laplacian and H1 gradients smoothly transform the initial geometry, the L2 projected gradient leads to more drastic shape and topology changes at the beginning of the optimization. This is not only due to the decreased regularity. The L2 projected gradient is not pulled down to zero in the vicinity of the forbidden region, thus allowing the breakwater to touch the forbidden region and to move away again in later iterations. If we introduce a rescaling of the gradient if it is larger than a given threshold the L2 projected gradient method is well behaved and actually often leads to better results than its smoother counterparts. This might be explained by a greater flexibility of the gradient which allows more local changes.

A note regarding topology changes: although the algorithm allows topology changes we do not compute a topological derivative. Hence, the sensitivity with regard to topology changes is not represented in the geometry update. This can lead to very small step sizes or even a stalling of the algorithm if this update leads to topology changes which increase the objective value. This observation motivates the use of non-monotone line search methods and is subject of current research.

Figure 2. Wave pattern for initial geometry of the first (left) and second experiment (right).

Figure 2. Wave pattern for initial geometry of the first (left) and second experiment (right).

Figure 3. Experiment 1: Convergence of the H1 projected gradient method (left) and the bi-Laplacian gradient method (right).

Figure 3. Experiment 1: Convergence of the H1 projected gradient method (left) and the bi-Laplacian gradient method (right).

Figure 4. Experiment 1: Wave pattern for the optimized geometry using the H1 projected gradient (left) and the bi-Laplacian gradient (right).

Figure 4. Experiment 1: Wave pattern for the optimized geometry using the H1 projected gradient (left) and the bi-Laplacian gradient (right).

Figure 5. Experiment 1: Wave pattern for the optimized geometry using the L2 projected gradient without rescaling (left) and with rescaling (right).

Figure 5. Experiment 1: Wave pattern for the optimized geometry using the L2 projected gradient without rescaling (left) and with rescaling (right).

Figure 6. Experiment 1: Convergence of the L2 projected gradient method without rescaling (left) and with rescaling (right).

Figure 6. Experiment 1: Convergence of the L2 projected gradient method without rescaling (left) and with rescaling (right).

The H1 gradient method reaches in iteration 1312 an L2 norm of the gradient below 1.0e-5 and is terminated. The objective value of 3.8e-7 is very close to the lower bound of 0. The bi-Laplacian gradient method gets stuck after 467 iterations. Whereas the L2 norm of the gradient 6.5e-5 is already pretty small the objective 2.6e-4 is several orders of magnitude higher than the result of the H1 gradient method. If we take a look at the convergence plots in Figure we see that in the beginning the bi-Laplacian gradient method reduces the L2 norm of the gradient much faster than the H1 gradient method. Comparing the final geometries in Figure shows that the lower regularity of the H1 gradient allows for more flexible changes which lead to a better objective value. The same behaviour could be observed in other tests as well. The bi-Laplacian gradient method reduces the L2 norm of the gradient quickly in few iterations, but then gets stuck. If we project the H1 gradient with respect to the L2 scalar product (effectively simply cutting it off in the forbidden regions) we observe a heavily deformed breakwater with topology changes, cf. Figure . But the convergence plot, cf. Figure , shows that the descent is very slow, after the first iterations only very small step sizes are chosen. In iteration 3000 the objective value is 7.4e-3 and the L2 norm of the gradient is 3.3e-2. If the mentioned rescaling of large gradients is introduced a much better performance can be observed. After 3000 iterations the objective value is 8.3e-6 and the L2 norm of the gradient is 2.5e-4.

Figure 7. Experiment 2: Wave pattern for the final geometry using the H1 projected gradient (left) and the bi-Laplacian gradient (right).

Figure 7. Experiment 2: Wave pattern for the final geometry using the H1 projected gradient (left) and the bi-Laplacian gradient (right).

Figure 8. Experiment 2: Convergence of the L2 projected gradient method without rescaling (left) and with rescaling (right).

Figure 8. Experiment 2: Convergence of the L2 projected gradient method without rescaling (left) and with rescaling (right).

Figure 9. Experiment 2: Wave pattern for the optimized geometry using the L2 projected gradient without rescaling (left) and with rescaling (right).

Figure 9. Experiment 2: Wave pattern for the optimized geometry using the L2 projected gradient without rescaling (left) and with rescaling (right).

In the second experiment the regular grid consists of 601×601 points and the incident wave direction is w=(0,-1)T. The initial geometry can be seen on the right side of Figure . After a good start the H1 projected gradient method is slowly decreasing the objective and after 3000 iterations achieves an objective value of 1.7e-5 and the L2 norm of the gradient is 4.7e-4. The bi-Laplacian gradient method gets stuck already after seven iterations. The final geometries are shown in Figure . In this example, the L2 projected gradient performs the best. After 3000 iterations the rescaled method achieves an objective value of 1.5e-6 and the L2 norm of the gradient is 8.4e-5. And without rescaling of the gradient it is even better. After choosing small step sizes in the first iterations the algorithm terminates in iteration 1337 with an L2 norm of the gradient below 1.0e-5 and an objective value of 6.3e-7. The convergence plots are shown in Figure and the optimized geometries in Figure .

In most of our experiments the volume of the breakwater did not vary much during the optimization. Still it might be interesting to add a volume constraint which might be realized by e.g. an Augmented Lagrange method.

9. Possible extensions and future research

This paper is just a first step towards a very interesting (real) application of shape optimization and should be considered a proof of concept. Various simplifications were made in the derivation of the physical model and need to be reconsidered in a more realistic setting. However, we think that these do not lead to fundamental new difficulties, only the derivation of the shape derivative and the actual implementation will require more effort. Furthermore, the model problem is easily extended to a situation where several different incident waves are considered. From the optimization point of view also various extensions are possible. As already mentioned the possibility of topology changes, which are not represented in the shape derivative, makes non-monotone line search methods a promising alternative. Moreover it would be interesting to consider, at least partially, second-order information in the scalar product (·,·)a or employ other speed-up strategies like non-linear conjugate gradients or approximate Newton-type methods. Finally, as mentioned in Section 5, other strategies for handling the geometric constraints are possible. Some of these topics are already subject of current research.

Acknowledgements

We wish to thank Prof. Wolfgang Ring (University of Graz) for valuable discussions and a careful proofreading of this article. Furthermore, we thank the anonymous referees for their feedback which helped us clarify the presentation.

Additional information

Funding

We gratefully acknowledge support from the International Research Training Group IGDK1754, funded by the German Science Foundation (DFG) and the Austrian Science Fund (FWF). The computations were carried out on a Linux cluster that was partially funded by the grant DFG INST 95/919-1 FUGG.

Notes

No potential conflict of interest was reported by the authors.

References

  • Allaire G, Jouve F, Toader A-M. A level-set method for shape optimization. C. R. Math. 2002;334:1125–1130. doi:10.1016/S1631-073X(02)02412-3. Available from: http://www.sciencedirect.com/science/article/pii/S1631073X02024123.
  • Allaire G, Jouve F, Toader A-M. Structural optimization using sensitivity analysis and a level-set method. J. Comput. Phys. 2004;194:363–393. doi: 10.1016/j.jcp.2003.09.032.
  • Osher SJ, Santosa F. Level set methods for optimization problems involving geometry and constraints: I. frequencies of a two-density inhomogeneous drum. J. Comput. Phys. 2001;171:272–288. doi: 10.1006/jcph.2001.6789. Available from: http://www.sciencedirect.com/science/article/pii/S0021999101967890.
  • Sethian J, Wiegmann A. Structural boundary design via level set and immersed interface methods. J. Comput. Phys. 2000;163:489–528. doi: 10.1006/jcph.2000.6581. Available from: http://www.sciencedirect.com/science/article/pii/S0021999100965811.
  • Wang MY, Wang X, Guo D. A level set method for structural topology optimization. Comput. Methods Appl. Mech. Eng. 2003;192:227–246. doi: 10.1016/S0045-7825(02)00559-5. Available from: http://www.sciencedirect.com/science/article/pii/S0045782502005595.
  • Allaire G, Dapogny C, Frey P. Topology and geometry optimization of elastic structures by exact deformation of simplicial mesh. C. R. Math. 2011;349:999–1003. doi: 10.1016/j.crma.2011.08.012. Available from: http://www.sciencedirect.com/science/article/pii/S1631073X1100241X.
  • Allaire G, Dapogny C, Frey P. A mesh evolution algorithm based on the level set method for geometry and topology optimization. Struct. Multi. Optim. 2013;48:711–715. doi: 10.1007/s00158-013-0929-2.
  • Allaire G, Dapogny C, Frey P. Shape optimization with a level set based mesh evolution method. Technical Report, Université Pierre et Marie Curie -- Paris VII; 2014.
  • Ha S-H, Cho S. Level set based topological shape optimization of geometrically nonlinear structures using unstructured mesh. Comput. Struct. 2008;86:1447–1455, structural Optimization. doi: 10.1016/j.compstruc.2007.05.025. Available from: http://www.sciencedirect.com/science/article/pii/S0045794907001757.
  • Persson P-O. Mesh generation for implicit geometries [PhD thesis]. Cambridge (MA): Massachusetts Institute of Technology; 2004.
  • Xia Q, Shi T, Liu S, Wang MY. A level set solution to the stress-based structural shape and topology optimization. Comput. Struct. 2012;90–91:55–64. doi: 10.1016/j.compstruc.2011.10.009. Available from: http://www.sciencedirect.com/science/article/pii/S0045794911002562.
  • Yamasaki S, Nomura T, Kawamoto A, Sato K, Nishiwaki S. A level set-based topology optimization method targeting metallic waveguide design problems. Int. J. Numer. Methods Eng. 2011;87:844–868. doi: 10.1002/nme.3135.
  • van Dijk N, Maute K, Langelaar M, van Keulen F. Level-set methods for structural topology optimization: a review. Struct. Multi. Optimiz. 2013;48:437–472. doi: 10.1007/s00158-013-0912-y.
  • Burger M, Osher SJ. A survey on level set methods for inverse problems and optimal design. Eur. J. Appl. Math. 2005;16:263–301. doi: 10.1017/S0956792505006182. Available from: http://journals.cambridge.org/articleS0956792505006182.
  • Allaire G. Shape optimization by the homogenization method. Vol. 58, Applied mathematical sciences. New York (NY): Springer; 2002.
  • Allaire G. Conception optimale de structures [Conception of optimal structures]. Vol. 58, Mathématiques & Applications. Berlin: Springer; 2007.
  • Bendsœ M. Optimization of structural topology, shape, and material. Berlin: Springer; 1995.
  • Delfour MC, Zolésio J-P. Shapes and geometries. 2nd ed. Philadelphia (PA): Society for Industrial and Applied Mathematics; 2011. doi: 10.1137/1.9780898719826. Available from: http://epubs.siam.org/doi/abs/10.1137/1.9780898719826.
  • Haslinger J, Mäkinen RAE. Introduction to shape optimization. Philadelphia (PA): Society for Industrial and Applied Mathematics; 2003. doi: 10.1137/1.9780898718690. Available from: http://epubs.siam.org/doi/abs/10.1137/1.9780898718690.
  • Henrot A, Pierre M. Variation et optimisation de formes. Une analyse géométrique [Variation and optimization of shapes. A geometric analysis]. Vol. 48, Mathématiques & Applications. Berlin: Springer; 2005.
  • Mohammadi B, Pironneau O. Applied shape optimization for fluids. Oxford: Oxford University Press; 2001.
  • Pironneau O. Optimal shape design for elliptic systems. Springer series in computational physics. Berlin: Springer; 1984.
  • Sokolowski J, Zolésio J-P. Introduction to shape optimization. Series in computational mathematic. Berlin: Springer; 1992.
  • Fernandes J, dos Santos MV, Fortes C. An element-by-element mild-slope model for wave propagation studies. In: ICS 2004 (Proceedings). J. Coastal Res. 2004:1869–1874.
  • Lee J-J, Lai C-P, Li Y. Application of computer modeling for harbor resonance studies of Long Beach & Los Angeles Harbor Basins. In: Coastal Engineering Proceedings, Vol. 1. Copenhagen; 1998.
  • Maa J-Y, Hwung H-H. A wave transformation model for harbor planning. In: Proceedings, Waves’97, Vol. 1. Virginia Beach, VA; 1997. p. 256–270.
  • Xing X. Computer modeling for wave oscillation problems in harbors and coastal regions [PhD thesis]. Los Angeles: University of Southern California; 2009.
  • Xing X, Lee J-J, Raichlen F. Harbor resonance: a comparison of field measurements to numerical results. In: Coastal Engineering Proceedings, Vol. 1. Shanghai; 2010.
  • Berkhoff J. Computation of combined refraction-diffraction. In: Proceedings 13th Coastal Engineering Conference. Vancouver: American Society of Civil Engineers; 1972. p. 471–490.
  • Baron Lopez FJ. Quelque problèmes d’optimisation de formes en électromagnétisme et méchanique de fluides [Some shape optimization problems in electromechanics and fluid mechanics][PhD thesis]. Paris VI, Grenoble, th. : analyse numéique; 1998.
  • Osher S, Sethian JA. Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys. 1988;79:12–49.
  • Fedkiw R, Osher S. Level set methods and dynamic implicit surfaces. Vol. 153, Applied mathematical sciences. New York (NY): Springer; 2003.
  • Giga Y. Surface evolution equations: a level set approach. Monographs in mathematics. Basel: Birkhäuser; 2006.
  • Sethian J. Level set methods and fast marching methods : evolving interfaces in computational geometry, fluid mechanics, computer vision, and materials science. 2nd ed. Cambridge (MA): Cambridge University Press; 1999.
  • Ihlenburg F. Finite element analysis of acoustic scattering. New York (NY): Springer; 1998.
  • Hiptmair R, Paganini A, Sargheini S. Comparison of approximate shape gradients. BIT Numer. Math. 2015;55:459–485. doi: 10.1007/s10543-014-0515-z.
  • Laurain A, Sturm K. Domain expression of the shape derivative and application to electricalimpedance tomography. Technical Report 1863, Weierstrass Institute for Applied Analysis and Stochastics; 2013.
  • Leis R. Initial boundary value problems in mathematical physics. Stuttgart: Wiley & Teubner Verlag; 1986.
  • Johnson C, Nedelec JC. On the coupling of boundary integral and finite element methods. Math. Comput. 1980;35:1063–1079. Available from: http://www.jstor.org/stable/2006375.
  • Rabinovich A. Seiches and harbor oscillations. In: Kim Y, editor. Handbook of coastal and ocean engineering. Singapore: World Scientific; 2009. p. 93–236.
  • Hackbusch W. Elliptic differential equations: theory and numerical treatment. Vol. 18, Springer series in computational mathematics; Berlin: Springer; 1992.
  • Delfour M, Zolésio J. Structure of shape derivatives for nonsmooth domains. J. Funct. Anal. 1992;104:1–33. doi: 10.1016/0022-1236(92)90087-Y. Available from: http://www.sciencedirect.com/science/article/pii/002212369290087Y.
  • Nečas J. Direct methods in the theory of elliptic equations. Berlin: Springer; 2012. doi: 10.1007/978-3-642-10455-8.
  • Hinze M, Pinnau R, Ulbrich M, Ulbrich S. Optimization with PDE constraints. New York (NY): Springer; 2009.
  • Blank L, Rupprecht C. An extension of the projected gradient method to a Banach space setting with application in structural topology optimization. Technical Report, ArXiv e-prints; 2015.
  • Ambrosio L, Crippa G. Existence, uniqueness, stability and differentiability properties of the flow associated to weakly differentiable vector fields. In: Ancona F, Bianchini S, Colombo RM, De Lellis C, Marson A, Montanari A. Transport equations and multi-D hyperbolic conservation laws. Vol. 5, Lecture Notes of the Unione Matematica Italiana. Berlin: Springer; 2008. p. 3–57. doi: 10.1007/978-3-540-76781-7.
  • Crandall MG, Hitoshi I, Lions PL. User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. 1992;27:1–67.
  • Osher S, Shu C. High-order essentially nonoscillatory schemes for Hamilton-Jacobi equations. SIAM J. Numer. Anal. 1991;28:907–922. doi: 10.1137/0728049. Available from: http://epubs.siam.org/doi/abs/10.1137/0728049.
  • Sethian JA. A fast marching level set method for monotonically advancing fronts. Proc. Nat. Acad. Sci. 1996;93:1591–1595.
  • Eaton JW, Bateman D, Hauberg S. GNU Octave version 3.0.1. manual: a high-level inter-active language for numerical computations. CreateSpace Independent Publishing Platform, 2009. ISBN 1441413006.
  • Kraft D. The “level-set” package for gnu octave octave forge, 2014. Available from: http://octave.sourceforge.net/level-set/.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.