Constrained Shape Derivative for the Spectral Laplace-Dirichlet Problem Using the Minimax Method ()
1. Introduction
In modern applied mathematics, more and more problems involve not just finding a function that satisfies an equation, but determining the very shape of the domain in which that function lives. These shapes are no longer fixed: they become variables in their own right, malleable and deformable, which we seek to adjust to achieve a given objective. Imagine a musical instrument, a drum for example. Its timbre depends closely on the shape of its membrane. Altering its contour, even slightly, changes its sound. Behind this acoustic intuition lies a profound mathematical truth: the eigenvalues of the Laplacian, which govern the frequencies of vibration, depend on the geometry of the domain. And if we want to optimise this geometry to obtain a deeper, higher or purer sound, we need to know how these eigenvalues change when we modify the shape of the domain. This is where the shape derivative comes in. There are several works in the literature on the shape derivative of the first eigenvalue. These include [1]-[6]. Here, we use other techniques to calculate the shape derivative of the first eigenvalue for the Laplace-Dirichlet problem. Next, we calculate the shape derivative of an objective function, taking the spectral problem as a constraint. Readers interested in the approach developed can consult the work of M. C. Delfour [7] [8] or our own [9] [10]. The shape derivative is a mathematical tool used to study the infinitesimal variation of a functional
when the domain is slightly deformed. This deformation is generally given by a vector field which defines a family of domains
, for
small or more generally
where the family of diffeomorphisms
is induced by the vector field
. The shape derivative of
is then defined by
This is the analogue of the directional derivative, but in the space of shapes (domains). The aim of this work is to calculate, using the min-max method [7], the shape derivative of the objective function
defined by
(1.1)
where
is the solution of the following eigenvalue problem
(1.2)
where
is a regular open domain of
,
a given vector of
and
a given function. The existence and uniqueness of this problem are detailed in [11]. Note that the problem under study has no source term. In other words,
is considered an eigenvalue and therefore depends on the domain
. The first step is to determine the derivative of this eigenvalue with respect to the domain. The rest of the manuscript is organized as follows: Section 2 presents a reminder of the methodology used to calculate the derivative. Section 3 is devoted to the study of the shape derivative of the volume function. Numerical simulations are proposed in Section 4, while the final section is devoted to results relating to the shape derivative in the context of the eigenvalue problem.
2. Methodology for Calculating Shape Derivatives
In this subsection, we describe how to calculate the topological derivative using the min-max approach, see e.g. [7] [9]. To begin with, we will look at the following definitions and notations.
Definition 2.1 A Lagrangian function is a function of the form
where X is a vector espace, Y a non empty subset of vector space and the function
is affine.
Associate with the parameter
the parametrized minimax
When the limits exist, we will use the following notations
The state equation at
(2.1)
The set of states
at
is denoted
(2.2)
The adjoint equation at
is
(2.3)
The set of solutions
at
is denoted
(2.4)
Finally the set of minimisers for the minimax is given by
(2.5)
Lemma 2.1 (Constrained infimum and minimax)
We have the following assertions
1)
2) The minimax
if and only if
. And in this case we have
.
3) If
, then
and
.
Proof. See [7]. ■
First Results of the Proposed Approach
We need the following assumption for everything that follows:
Hypothesis (H0)
Let X be a vector space.
1) For all
,
,
and
, the function
is absolutely continuous. This implies that for almost all
the derivative exists and is equal to
and it is the integral of its derivative. In particular
2) For all
,
,
and
,
and for almost all
,
exist et the functions
belong to
Definition 2.2 Given
and
, the averaged adjoint equation is:
and the set of solutions is noted
.
clearly reduces to the set of standard adjoint states
at
.
Theorem 2.2 Consider the Lagrangian functional
where X and Y are vector spaces and the function
is affine. Assume that (H0) and the following hypotheses are satisfied
(H1) for all
,
is finite,
and
are singletons,
(H2)
exists,
(H3) The following limit exists
Then,
exists and
.
Proof. Recall that
and
for each
, then for a standard adjoint state
at
Dividing by
, we obtain
Going to the limit when
goes to zero, we obtain
■
Corollary 2.3 Consider the Lagrangian functional
where X and Y are vector spaces and the function
is affine. Assume that (H0) and the following assumptions are satisfied:
(H1a) for all
,
,
is finite, and for each
,
,
(H2a) for all
and
exists,
(H3a) there exist
and
such that the following limit exists
Then,
exists and there exist
and
such that
.
Proposition 2.4 Let X and Y be two vector spaces and
be a Lagrangian. If the function
is affine, then for all
, we have
Proof. The Lagrangian
is affine in
means
, où
is linear in
.
Since
does not depend on
, we have
By definition, the following applies
So for all
(2.6)
and it follows that
(2.7)
Now let us calculate
.
But
because
is linear in
and it follows
(2.8)
So the relationships (2.6), (2.7) and (2.8) give the result. ■
3. Shape Derivative for the Volume Function via Min-Max Approach
The following example is a special case where we consider a shape functional independent of a PDE constraint. It is given in [5] by
(3.1)
is the volume of
et
is the first eigenvalue of the following eigenvalue problem
(3.2)
This problem has a sequence of real positive eigenvalues:
with associated eigenfunctions
, orthonormal in
, i.e.
In this case, we need the volume expression of the derivative of the eigenvalue. For this we have the following proposition.
Proposition 3.1 Let
be a bounded open domain. Assume that the first eigenvalue
is simple. Then
is shape differentiable and the Eulerian derivative is
(3.3)
If, in addition,
is convex or of class
, then
(3.4)
Proof. A proof of this proposition can also be found in the work of A. Henrot and M. Pierre [4] or that of S. Zhu [5]. We only prove the derivative formula in volume form (3.3). The reader can also consult [6]. In this proof, we will therefore use a rigorous demonstration based on the min-max method introduced by M. C. Delfour in [7]. To do this, we consider the diffeomorphism family
and define the perturbed domain by
. The perturbed state
is the solution of the variational problem
A change of variables allows the problem to be transported to the reference domain
. Thus, noting by
and
,
is the solution of the following variational equation
(3.5)
For
(3.6)
Deriving with respect to
and taking
, we obtain
(3.7)
So at
with
.
(3.8)
Using the normalization condition, we have
(3.9)
with
. Using the fact that
(3.10)
■
Proposition 3.2 Let
be the objective function defined by
. Then the derivative of
in
in the direction
is given by
(3.11)
Proof. If we consider the objective function
(3.12)
first, we know that the derivative of the volume functional is given by
Indeed, by definition of the shape derivative
We use a change of variables to transform the domain
into
Since
and
, we have well
. And finally, the derivative of the objective function
is given by
■
4. Numerical Tests
In shape optimization, the aim is to find the most efficient shape for a certain criterion in an admissible set of domains. In this section, we present numerical tests for minimizing the objective function defined in (3.1) taking into account that
is the first eigenvalue of the operator
.
Shape Optimization Implementation Details
The interest of the shape derivative is to give a direction of descent to evolve the domain in order to minimize the eigenvalue. The idea is to find a vector field V that gives a decrease in the objective function, i.e. we want to find a vector field V such that
. This can be achieved by solving an auxiliary limit problem of the type
for a suitable Herbert space
.
is a positive bilinear form that we are free to choose. Then
is a descent direction because
We present a shape gradient descent algorithm, based on a volume expression rather than a boundary expression. In this case, it will be essential to update the finite element mesh after each iteration. To this end, we follow the approaches developed by V. H. Schulz, M. Siebenborn, K. Welker, [12] using the linear elasticity equation given as follows
where
and
are the deformation and continuity tensors,
is the identity tensor,
is the trace operator on a tensor and
is the displacement vector field [12] [13].
and
designate the Lamé parameters, which can be expressed in terms of Young's modulus
and Poisson’s ratio
as follows
The Young’s modulus
and Poisson’s ratio
have effects on mesh deformation.
indicates the stiffness of the material, allowing us to control the step size for shape updating, while
gives the ratio between mesh expansion in other coordinate directions and compression in a particular direction. These coefficients can be found, for instance, in [12] [13]. The bilinear form of the linear elasticity equation is given by
Thanks to the Riesz representation theorem, we obtain the direction of descent or the shape gradient or the solution of the linear elasticity equation
, which can be solved as follows
The shape functional is not subject to any PDE constraints. It is therefore sufficient to solve the state equation and the deformation equation at each iteration step using the following algorithm:
Algorithm 1 |
Initialization while
do 1. Calculate the solution
of the spectral problem. 2. Calculate the gradient
[via
and Linear elasticity]. 3. Deforming the domain
in
[via ALE.move and
]. end while |
First, it is worth noting that all numerical simulations are conducted in two dimensions within the FEniCS framework, see for instance [13] to find out more. The algorithm used here is a slight modification of the one proposed in [6], in which
is assigned a value. In our case, since
is an eigenvalue, it is calculated directly. In the absence of a partial differential equation constraint, each iteration consists of solving the equation of state and the system associated with the direction of descent, or simply the linear elasticity equation. As we shall see, obtaining the new domain shape relies solely on moving the points on the boundary. We implement the general problem of minimizing the functional (3.1) depending on
, eigenvalue for (3.2). For each iteration, we need to solve the state equation (3.2). The shape derivative we use is given by the formula (3.11). To define the Lame parameters, we fixed the values
and
throughout the numerical tests. We set the time step to 0.001. We follow the Algorithm 1 to obtain a decrease in the functional introduced in (3.1). Tests were first carried out for 100 iterations, then extended to 500 and finally 1000 iterations. The convergence of the algorithm depends on the tolerance chosen which is set here at 0.088. The initial mesh is the unit square, as shown in Figure 1.
After some numerical testing, the vector fields look like this: Figure 2 shows the vector fields for 100, 500 and 1000 iterations.
Figure 3, Figure 4 and Figure 5 show the optimized domains, the solution of the state equation and the evolution of the objective function for 100, 500 and 1000 iterations respectively.
Figure 1. Initial mesh.
Figure 2. Vectors fields after 100, 500, and 1000 iterations.
Figure 3. (1): Optimized domain, (2): State, (3): Objective function for 100 iterations.
Figure 4. (1): Optimized domain, (2): State, (3): Objective function for 500 iterations.
Figure 5. (1): Optimized domain, (2): State, (3): Objective function for 1000 iterations.
We also carried out tests with the unit disk. Given the slow rate of convergence, the number of iterations was limited to 500. The simulation results are presented in Figure 6 below.
Figure 6. (1): Optimized domain, (2): State, (3): Objective function for 500 iterations.
5. Shape Derivative of an Eigenvalue Problem Using a Min-Max Approach
In this section, we focus on calculating the shape derivative in the presence of a spectral problem as a constraint. We use the same computational techniques developed in Section 2. The difference is that here, the domain-dependent
is considered as the first eigenvalue of the Dirichlet Laplacian.
Let
be a bounded lipschitzian open domain in
with a regular boundary
. Let
be the solution to the following spectral problem
(5.1)
is considered here as the first eigenvalue of the Laplacian. We consider the functional
defined by
(5.2)
where
is the solution of (5.1), A a given vector of
and
a given function. As in the case of the HeLmholtz equation [9], we want to calculate the shape derivative of the function
. Recall that the shape derivative of
in
in the direction
is defined by
when the limit exists where
. We will also follow M. C. Delfour [7] [8] to calculate the shape derivative of the
functional. The functional in the domain
is then defined by
(5.3)
where
is the solution to the problem
(5.4)
is the boundary of
. The variational formulation of (5.4) is to find
such that
(5.5)
Note that this problem is not trivial, as the operator acts on domains that change
. The eigenfunctions
live in different spaces
. The normalization
also depends on
. To solve this problem, we transport the domain onto a fixed
domain, via the change of variables
. This will enable us to manipulate objects in a fixed frame. To this end, we introduce the following parameterization
(5.6)
and to work in
, we also introduce the compositions
,
and
. So the variational formulation of (5.4) amounts to looking for
such that
(5.7)
Applying the change of variables formula, we obtain
In addition, we have
Let
be the Jacobian matrix of
evaluated in
,
the transpose matrix of
, then we have the following proposition
Proposition 5.1 We have
1)
2)
Proof. See [1]. ■
So using the proposition, we can write
Finally, the previous equation becomes
(5.8)
with
,
,
the Jacobian matrix of
.
To write the functional in the fixed domain
, we first assume that A is written
and note by
.
Applying Proposition 5.1 once more, we obtain
(5.9)
Since the set of perturbed states admits a solution, then
. So the
-dependent Lagrangian can be defined as follows:
Let
be the following function
We assume that the function
admits a finite limit noted
, then we have the following proposition.
Proposition 5.2 Let
be the solution to the problem
, then the shape derivative of the functional
defined in (5.2) is given by
where
solution of (5.1) and
, the adjoint state solution of
Proof. The derivative of the
-dependent Lagrangian with respect to
.
Taking
and
, we obtain
Taking
equal to zero, we get
The derivative of the
-dependent Lagrangian with respect to
in the direction
:
The derivative of the
-dependent Lagrangian with respect to
in the direction
:
(5.10)
The state equation at
and the adjoint state equation at
are respectively given by
(5.11)
(5.12)
Taking
in adjoint Equation (5.12), we have
Consequently, we get
In other words,
Accordingly, the function
can be rewritten as follows:
As in the case of the Helmholtz equation studied in [9], we have the following inequality
Now we need to show that the limit of
exists. First, we know the terms
and
are uniformly bounded. It will therefore suffice to show that
in
-strong and that the
-norm of
is bounded. But the difficulty lies in the fact that the eigenvalue
depends on the variable
and we have no information on whether
is bounded. So we just assume that
admits a finite limit denoted
and in this case the shape derivative of the functional
is given by:
where
solution of (5.1) and
, the adjoint state solution of
■
6. Conclusion
This work provides a simple method for calculating the shape derivative in the context of Dirichlet eigenvalue problems. We have used alternative techniques to calculate the shape derivative of the first eigenvalue of the Laplace-Dirichlet problem, as well as that of an objective function constrained by a spectral problem. A fundamental tool in shape optimization, the shape derivative enables us to analyze how a quantity of interest varies under the effect of small deformations of the domain. This approach, inspired in particular by the work of Delfour, opens the way to a better understanding and control of geometry-dependent spectral problems. Some numerical tests were also carried out in two dimensions. The next step is to calculate the topological derivative using the same approach for an eigenvalue problem.