Next Article in Journal
On One Approach to Obtaining Estimates of the Rate of Convergence to the Limiting Regime of Markov Chains
Previous Article in Journal
Class-Aware Self- and Cross-Attention Network for Few-Shot Semantic Segmentation of Remote Sensing Images
Previous Article in Special Issue
Non-Invasive Identification of Vehicle Suspension Parameters: A Methodology Based on Synthetic Data Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Concept of Topological Derivative for Eigenvalue Optimization Problem for Plane Structures

by
Fernando Soares Carvalho
1,* and
Carla Tatiana Mota Anflor
2,†
1
Department of Mathematics, Federal University of Tocantins, Palmas 77650-000, Brazil
2
Group of Experimental and Computational Mechanics, University of Brasilia, Brasilia 70910-900, Brazil
*
Author to whom correspondence should be addressed.
Arraias 77330-000, Brazil.
Mathematics 2024, 12(17), 2762; https://doi.org/10.3390/math12172762
Submission received: 18 June 2024 / Revised: 30 August 2024 / Accepted: 3 September 2024 / Published: 6 September 2024

Abstract

:
This paper presents the topological derivative of the first eigenvalue for the free vibration model of plane structures. We conduct a topological asymptotic analysis to account for perturbations in the domain caused by inserting a small inclusion. The paper includes a rigorous derivation of the topological derivative for the eigenvalue problem along with a proof of its existence. Additionally, we provide numerical examples that illustrate the application of the proposed methodology for maximizing the first eigenvalue in plane structures. The results demonstrate that multiple eigenvalues were not encountered.

1. Introduction

Some considerations such as dynamic response and loads are important for understanding the behavior and design of structures. Structures under dynamic analysis are governed by linear differential equations, which involve solving an eigenvalue problem. The eigenvalue problem is crucial in many real-world applications, particularly in structural optimization, as it significantly contributes to structural integrity. When the excitation frequency matches one of the natural frequencies of vibration, the structure becomes highly responsive, which can lead to damage or collapse. Structural behavior is affected when variations in system parameters lead to changes in eigenvalues; eigenvectors; and, consequently, the final response characteristics of the system. The magnitude of these variations is reflected in the derivatives of the system’s eigenvalues and eigenvectors. One of the main concerns of sensitivity analysis is the presence of multiple eigenvalues. During the topology optimization process, only simple eigenvalues are present in the initial steps. However, as the iterative process progresses and the geometry changes, multiple eigenvalues may appear. The presence of multiple eigenvalues leads to convergence issues in optimization algorithms because they are not differentiable in the common sense [1,2]. Since then, several methods have been proposed to avoid the presence of multiple eigenvalues [3,4,5,6]. The conventional notion of derivative is naturally extended to functionals through the concept of the topological derivative ( D T ), where the variable is a geometric domain subject to singular topology changes.
Concerning structural topology optimization, the D T provides the exact sensitivity of the associated objective functionals due to perturbations such as the insertion of infinitesimal voids, inclusions, or even cracks. In particular, for dynamic problems, the D T for simple eigenvalues of the Laplacian was considered by [7], and for multiple eigenvalues in elasticity problems by [8]. A similar work concerning the augmented Lagrangian method based on the concept of the topological derivative was also addressed by [9], where numerical results considering compliance and eigenfrequency constraints were introduced by the authors. More recently, the D T of L 2 and energy norms associated with the solution to plate bending, considering both Kirchhoff and Reissner–Mindlin theories, was introduced by [10]. In addition to the maximization of the first eigenvalue, the sensitivities obtained were adapted to the context of plate topology optimization for plates under elastic support and free vibration conditions. In [11], the D T was derived for membrane eigenvalue maximization concerning the nucleation of inclusions endowed with different material properties from the background. According to the numerical results, the presence of multiple eigenvalues was not observed as the iterative process evolved, even as the geometry became complex.
The field of eigenvalue topology optimization remains significant and requires the development of new methodologies. These should incorporate classical techniques such as sensitivity analysis, level set methods, and the bubble method, among others, to effectively address various challenges. Ref. [12] introduced a phase field-based structural optimization method that simplifies computation and improves efficiency compared to traditional level-set methods. This method uses a phase field function and solves a time-dependent reaction–diffusion equation but relies on the initial shape and requires careful management of perimeter control effects. More recently, ref. [13] presented an enhanced phase field method with multi-level correction, which improves accuracy and reduces computational costs for eigenfrequency topology optimization, as demonstrated by numerical examples in 2D and 3D.
As mentioned earlier, eigenfrequency optimization is a prominent research topic. The present work revisits the subject discussed in [11] to incorporate the problem of plane structures. The focus is on deriving the D T for maximizing the first eigenvalue. A closed formula for the D T of the first eigenvalue for plane structures is obtained. The resulting formula is used together with a level-set domain representation method to develop a topology design algorithm, as implemented in [14]. The optimal topology for the plane structure under prescribed boundary conditions is achieved by maximizing the first eigenvalue while minimizing the volume. Several proposed numerical examples demonstrate the feasibility of the present methodology in addressing topology optimizations. It is worth mentioning that, to the authors’ knowledge, the concept of D T for the eigenvalue problem for plane structures, considering mathematical rigor, has not been addressed elsewhere. Additionally, several numerical examples are explored. This paper is organized as follows. The topological derivative for plane structure problems is introduced in Section 2. In Section 3, some numerical examples are presented, showing that the D T derived here successfully solves eigenvalue problems for plane structures. Finally, the conclusions are summarized in Section 4.

2. Topological Derivative

Consider, for instance, an open and bounded domain Ω R 2 such that it is subject to a nonsmooth perturbation confined in a small ball B ε ( x ^ ) of radius ε and center at x ^ Ω ,
ψ ( χ ε ( x ^ ) ) = ψ ( χ ) + f ( ε ) D T ψ ( x ^ ) + o ( f ( ε ) ) ,
where ψ ( χ ) is the shape functional associated with the unperturbed domain, f ( ε ) is a positive first-order correction function of ψ , and o ( f ( ε ) ) is the remainder, such that o ( f ( ε ) ) / f ( ε ) 0 as ε 0 . The function χ is the characteristic function associated with the unperturbed domain, and χ ε is the characteristic function associated with the perturbed domain. The function x ^ D T ψ ( x ^ ) is termed the topological derivative of ψ at x ^ . The domain Ω is then divided into two subdomains, ω Ω and its complement Ω ω . Finally, a set of piecewise constant functions α , ρ , and β (which are the contrasts in the material properties) is considered and introduced according to Table 1.
The topological perturbation results from the nucleation of a small circular inclusion of the form ω ε ( x ^ ) : = B ε ( x ^ ) = | x x ^ | < ε for x ^ Ω . In this specific case, the perturbation is governed by a set of piecewise constant functions α ε , ρ ε , and β ε , as introduced in Table 2 and Table 3.

2.1. Plane Structures

This section introduces the mathematical model for the plane structure problem and the shape functionals previously introduced herein in relation to the eigenvalue problem, for the sake of completeness. It also covers both the perturbed and unperturbed problems and demonstrates the existence of the associated topological derivative.
The original unperturbed problem can be stated as follows: Find u V ( Ω ) , such that
Ω α σ ( u ) · s v + Ω ρ k u · v = Ω β f · v v V ( Ω ) ,
where V ( Ω ) = H 0 1 ( Ω ; R 2 ) . The coefficients α , ρ , and β R + , are given in Table 1. In addition, σ ( u ) = C s u , u : Ω R 2 is the displacement field and k is a positive function. The symmetric part of the displacement gradient tensor s u and constitutive tensor C are given by
s u = u + ( u ) T 2 and C = E 1 + ν I + ν 1 ν I I ,
where the symbols I and I represent the second- and fourth-order identity tensors, respectively. It is worth mentioning that only the homogeneous Dirichlet condition on the boundary Ω was considered. However, this could be replaced by any other boundary condition, provided the problem of interest remains well-posed.
Additionally, ν is the Poisson ratio and E is the Young’s modulus. The L 2 and energy norm shape functionals we are dealing with are defined as follows:
G ( u ) = Ω ρ k u 2 and J ( u ) = Ω α σ ( u ) · s u .
To simplify the expression of the topological derivatives, we introduce the adjoint problems for displacements q and p as follows:
q V ( Ω ) : Ω α σ ( q ) · s v + Ω ρ k q · v = 2 Ω ρ k u · v , v V ( Ω ) ,
p V ( Ω ) : Ω α σ ( p ) · s v + Ω ρ k p · v = 2 Ω α σ ( u ) · s v , v V ( Ω ) .
The topologically perturbed counterpart of problem (2) is expressed as follows: Find u ε V ( Ω ) , such that
Ω α ε σ ( u ε ) · s v + Ω ρ ε k u ε · v = Ω β ε f · v v V ( Ω ) ,
where the coefficients α ε , ρ ε , and β ε are defined through Table 2 and Table 3. The associated shape functionals are then defined as
G ε ( u ε ) = Ω ρ ε k u ε 2 and J ε ( u ε ) = Ω α ε σ ( u ε ) · s u ε .

2.1.1. Existence of the Topological Derivative

Equations (2) and (7) introduce the shape functionals for the original and perturbed domains. Based on these, the existence of the associated topological derivative can be stated as follows:
Lemma 1.
Let u and u ε be solutions to the original (2) and perturbed (7) problems, respectively. Then, the estimate u ε u H 1 ( Ω ; R 2 ) = O ( ε ) holds true.
Proof. 
Let us subtract (2) from (7). By setting v = u ε u as the test function, after some simple analytical work, there is
Ω α ε σ ( u ε u ) · s ( u ε u ) + Ω ρ ε k u ε u 2 = B ε ( 1 γ α ) α σ ( u ) · s ( u ε u ) + B ε ( 1 γ ρ ) ρ k u · ( u ε u ) B ε ( 1 γ β ) β f · ( u ε u ) ,
where the contrasts introduced in Table 2 and Table 3 are considered. The Cauchy–Schwarz inequality yields
Ω α ε σ ( u ε u ) · s ( u ε u ) + Ω ρ ε k u ε u 2 C 1 ε u ε u H 1 ( Ω ; R 2 ) ,
where the elliptic regularity of function u is used. The coercivity of the bilinear form on the left-hand side of Equation (10) results
c u ε u H 1 ( Ω ; R 2 ) 2 Ω α ε σ ( u ε u ) · s ( u ε u ) + Ω ρ ε k u ε u 2
which leads to the result
u ε u H 1 ( Ω ; R 2 ) C ε ,
with the constant C = C 1 / c independent of the small parameter ε . □

2.1.2. Topological Sensitivities

The following fourth-order polarization tensor associated with the elasticity model is introduced as
P = 1 2 1 γ α 1 + γ α δ 2 ( 1 δ 2 ) I + 1 2 ( δ 1 δ 2 ) 1 γ α 1 + γ α δ 1 I I .
The constants δ 1 and δ 2 are given by
δ 1 = μ l + λ l μ l , δ 2 = 3 μ l + λ l μ l + λ l ,
where μ l and λ l are Lame’s coefficients, both considered constants everywhere. In this case, the plane stress assumption is applied as
μ l = E 2 ( 1 + ν ) λ l = ν E 1 ν 2 .
The two main results of this section are stated here and are analogous to the work of [11]. The proofs of Theorems 1 and 2 are in the Appendix A.
Theorem 1.
Let G ( u ) be the shape functional defined by (4)-left; then, its associated topological derivative is given by
D T G = α P σ ( u ) · s q ( 1 γ ρ ) ρ k u · ( u + q ) + ( 1 γ β ) β f · q a . e . in Ω
where q is the adjoint state solution of (5).
Theorem 2.
Let J ( u ) be the shape functional presented in (4)-right; then, its topological derivative is given by
D T J = α P σ ( u ) · s ( u + p ) ( 1 γ ρ ) ρ k u · p + ( 1 γ β ) β f · p a . e . in Ω
where p is the adjoint solution of problem (6).

2.2. Eigenvalue Problem

The eigenvalue problem for the plane structures model of clamped free vibration can be stated as follows: Find u and λ , such that
div ( α σ ( u ) ) = λ ρ u in Ω , u = 0 on Ω .
One can define the associated first eigenvalue as
λ 1 = inf u H 0 1 ( Ω ; R 2 ) , u 0 Ω α σ ( u ) · s u Ω ρ u 2 ,
with u being the solution of (18).
Note that D T obeys the quotient rule for differentiation and uses the functions described in Theorems 1 and 2. In particular, since the topological derivative obeys the basic rules of the differential calculus (including the quotient rule for differentiation), the rigorous justification for these kind of results can be found in the book by [15] (Ch 9). The topological derivative of
J ( D ) : = λ 1 ,
is given by
D T J = α P σ ( u ) · s u + ( 1 γ ρ ) ρ λ 1 u 2 Ω ρ u 2 .

3. Numerical Results

In this section, several numerical examples are presented to demonstrate the capability of the topological derivative concept in delivering optimal topologies for the problems addressed here. Based on the results from the numerical examples presented, the effectiveness of the proposed methodology for optimizing the first eigenvalue under various boundary conditions and loading scenarios will be evaluated.
In all examples, the topological optimization problem focuses on maximizing the first eigenvalue while adhering to a volume constraint. The convergence to a local minimum is demonstrated through various numerical examples. We apply a topology optimization algorithm that employs an evolution equation for the level-set function, based on D T (for further details, see [14]).

First Eigenvalue Maximization

Three cases are presented: Cases A, B, and C. We introduce a hold-all domain D R 2 such that Ω D . The hold-all domain is discretized using linear triangular finite elements, resulting in an initial grid with uniform mesh spacing.
In Case A, the hold-all domain D is a rectangle ( 0 , 3 ) × ( 0 , 1 ) , m 2 , clamped on the left side along the contours Γ D 1 = 0 × ( 0 , 0.1 ) and Γ D 2 = 0 × ( 0.9 , 1.0 ) , with a concentrated mass m = 0.4 located at the point ( 3 , 0.5 ) , as depicted in Figure 1a. Figure 1b shows the displacement in the initial domain.
In Case B, the hold-all domain D is a rectangle ( 0 , 6 ) × ( 0 , 1 ) , m 2 , clamped at the point ( 0 , 0 ) and simply supported at the point ( 6 , 0 ) . The concentrated masses m = 0.7 are located at the points ( 1 , 1 ) and ( 5 , 1 ) , and an additional concentrated mass m 1 = 1.4 is located at the coordinate ( 3 , 1 ) , as depicted in Figure 2a. The displacement is shown in Figure 2b.
In Case C, the hold-all domain D is a rectangle ( 0 , 4 ) × ( 0 , 1 ) , m 2 . Concentrated masses m = m i , where i = 1 , 2 , , 11 with m i = 70 , are uniformly distributed along the upper side of the rectangle (see Figure 3a). Figure 3b shows the displacement in the initial domain.
Linear triangular finite elements were employed for discretization in all the cases presented.
Young’s modulus is set as E = 210 GPa and contrast parameters as γ α = γ ρ = 10 3 . Information on the initial grid, number of elements and number of nodes in the initial mesh, Poisson’s ratio ν , and penalty parameter μ are expressed in Table 4, for each case.
It should be noted that mesh refinement procedures were employed as part of the optimization scheme to improve the boundary resolution of the final topology. Table 4 provides details on the final mesh resulting from the refinement process.
To enhance accuracy and achieve smoother topologies, four levels of uniform mesh refinement were implemented during the iterative process for Cases A and B, and five levels for Case C.
The mesh refinement levels resulted in 2,764,600 elements and 1,384,321 nodes for Case A, 5,529,600 elements and 2,768,161 nodes for Case B, and 409,600 elements and 205,601 nodes for Case C.
The optimal topologies for Cases A, B, and C are illustrated in Figure 4, Figure 5, Figure 6 and Figure 7, respectively. The effect of mesh refinement on the optimized topology for Case A is evident (see Figure 5). The final topology for Case B (Figure 6) exhibited asymmetry due to the boundary conditions, which included clamping at point (0,0) and simple support at point (6,0). In this example, the initial mesh comprises 19,200 elements (40 × 120 grid) and 9761 nodes, with all other parameters and boundary conditions consistent with those described for Case A.
Figure 8 introduces the normalized first eigenvalue history λ 1 / λ 1 0 as the iterative process has evolved while Figure 9 depicts the normalized first eigenvalues’ history λ 1 / λ 2 .
It can be observed that Case C achieved a higher ratio of λ 1 / λ 2 compared to Cases A and B.
However, during the optimization process, no coincident eigenvalues were observed. Figure 10 and Figure 11 illustrate the evolution of the shape functional and volume, respectively.
To evaluate the influence of the parameter μ on the final design, a set of four cases is considered. Variations in the parameter μ are detailed in Table 5.
For Cases D1, D2, D3, and D4, the hold-all domain D is a unit square ( 0 , 1 ) × ( 0 , 1 ) , m 2 , clamped at the points ( 0 , 0 ) and ( 0 , 1 ) . A concentrated mass m = 0.1 is applied at the point ( 1 , 1 ) , as shown in Figure 12a, with the initial displacement depicted in Figure 12b.
The final topologies for each case are presented in Figure 13a–d. Figure 14 and Figure 15 show the history of the shape functional and volume fraction, respectively.
The history of the normalized first eigenvalue, λ 1 / λ 1 0 (where λ 1 0 is its initial value), is illustrated in Figure 16, showing the evolution until the process is halted.
Figure 17 depicts the history of the normalized first eigenvalue ratio, λ 1 / λ 2 . As before, no presence of multiple eigenvalues was observed.

4. Conclusions

In this work, the associated topological derivative of the first eigenvalue for plane structure problems was used for topology optimization. The convergence to a local minimum was achieved by combining the level set method with the analytical formula for the topological derivative ( D T ). While coincident eigenvalues are often encountered in this type of problem, they were not observed in the present work. The algorithm converged successfully for all numerical examples, and no special methods were needed to handle multiple eigenvalues. The results highlight the effectiveness of the D T concept in deriving optimal topologies for eigenvalue problems in plane structures.

Author Contributions

Writing—original draft, F.S.C. and C.T.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Some or all data or code generated or used during the study are available from the corresponding author by reasonable request (ETABS command lines files and MATLAB R2022a code used to perform analyses).

Acknowledgments

The authors gratefully acknowledge the valuable contributions given by A.A. Novotny during the topological derivative mathematical development. Additionally, the authors also extend their thanks to S.M. Giusti for kindly sharing the numerical routine, which was adapted for the present topology optimization procedure. Finally, we would like to express our gratitude to Dirley Rucheinsky (in memoriam) for his valuable contributions to the numerical results presented in this work.

Conflicts of Interest

No potential conflicts of interest were reported by the authors.

Appendix A. Topological Asymptotic Analysis

Let us introduce an ansatz for the solution u ε to the perturbed boundary value problem (7) of the form
u ε ( x ) = u ( x ) + w ε ( x ) + u ˜ ε ( x ) ,
where u is the solution to the unperturbed boundary value problem (2), w ε is the solution to an exterior boundary value problem, and u ˜ ε is the remainder.
In particular, the exterior problem reads as follows: Find w ε , such that
div ( α ε σ ( w ε ) ) = 0 in R 2 , w ε 0 at ( w ε ) ( α ε σ ( w ε ) ) · n = = 0 g on B ε ,
where g = ( 1 γ α ) α σ u ( x ^ ) · n . The above boundary value problem admits an explicit solution (see, for instance, the book by Novotny and Sokolowski [16]), which can be written in a polar coordinate system ( r , θ ) with a center at x ^ , as follows:
For r ε (outside the inclusion),
σ r r ( w ε ( r , θ ) ) = φ 1 1 γ α 1 + γ α a 1 ε 2 r 2 φ 2 4 1 γ α 1 + γ α a 2 ε 2 r 2 + 3 1 γ α 1 + γ α a 2 ε 4 r 4 c o s 2 θ ,
σ θ θ ( w ε ( r , θ ) ) = φ 1 1 γ α 1 + γ α a 1 ε 2 r 2 φ 2 3 1 γ α 1 + γ α a 2 ε 4 r 4 c o s 2 θ ,
σ r θ = φ 2 2 1 γ α 1 + γ α a 2 ε 2 r 2 3 1 γ α 1 + γ α a 2 ε 4 r 4 s i n 2 θ .
For 0 < r < ε (inside the inclusion),
σ r r ( w ε ( r , θ ) ) = φ 1 a 1 γ α 1 γ α 1 + γ α a 1 + φ 2 a 2 γ α 1 γ α 1 + γ α a 2 c o s 2 θ ,
σ θ θ ( w ε ( r , θ ) ) = φ 1 a 1 γ α 1 γ α 1 + γ α a 1 φ 2 a 2 γ α 1 γ α 1 + γ α a 2 c o s 2 θ ,
σ r θ ( w ε ( r , θ ) ) = φ 2 a 2 γ α 1 γ α 1 + γ α a 2 s i n 2 θ .
Some terms in the above formulae require explanations. The coefficients φ 1 and φ 2 are given by
φ 1 = 1 2 ( σ 1 ( u ( x ^ ) ) + σ 2 ( u ( x ^ ) ) ) , φ 1 = 1 2 ( σ 1 ( u ( x ^ ) ) σ 2 ( u ( x ^ ) ) ) ,
where σ 1 ( u ( x ^ ) ) and σ 2 ( u ( x ^ ) ) are the eigenvalues of tensor σ ( u ( x ^ ) ) , which can be expressed as
σ 1 , 2 ( u ( x ^ ) ) = 1 2 t r σ ( u ( x ^ ) ) ± 2 σ D ( u ( x ^ ) ) · σ D ( u ( x ^ ) ) ,
with σ D ( u ( x ^ ) ) standing for the deviatory part of the stress tensor σ ( u ( x ^ ) ) , namely,
σ D ( u ( x ^ ) ) = σ ( u ( x ^ ) ) 1 2 t r σ ( u ( x ^ ) ) I .
In addition, the constants a 1 and a 2 are given by
a 1 = μ + λ μ , 3 μ + λ μ + λ .
Finally, σ r r ( u ε ) , σ θ θ ( u ε ) , and σ r θ ( u ε ) are the components of tensor σ ( u ε ) in the polar coordinate system, namely, σ r r ( u ε ) = e r · σ ( u ε ) e r , σ θ θ ( u ε ) = e θ · σ ( u ε ) e θ , and σ r θ ( u ε ) = e r · σ ( u ε ) e θ , with e r and e θ used to denote the canonical basis associated with the polar coordinate system ( r , θ ) ), such that e r = e θ = 1 and e r · e θ = 0 (for more details, see [16]).
Note that the remainder is constructed in order to compensate for the discrepancies introduced by the boundary layers w ε and by the higher-order terms of the Taylor series expansion of σ ( u ) around the point x ^ Ω . It means that u ˜ ε has to be the solution to the following boundary value problem: Find u ˜ ε , such that
div [ α ε σ ( u ˜ ε ) ( x ) ] + ρ ε k u ˜ ε ( x ) = ρ ε k w ε in Ω , u ˜ ε = ε 2 g 1 on Ω , ( u ˜ ε ) ( α ε σ ( u ˜ ε ) ) n = = 0 ε g 2 on B ε ,
with functions g 1 = ε 2 w ε and g 2 = ( 1 γ α ) α [ σ ( w ε ] n · n independent of ε .
Lemma A1.
Let u ˜ ε be the solution of (A13) or, equivalently, of the following variational problem
u ˜ ε U ˜ ε : Ω α ε σ ( u ˜ ε ) · v + Ω ρ ε k u ˜ ε v + = Ω ρ ε k w ε v + ε B ε g 2 v , v H 0 1 ( Ω ) ,
where the set U ˜ ε : = { φ H 1 ( Ω ) : φ | Ω = ε 2 g 1 } . Then, the estimate u ˜ ε H 1 ( Ω ) = o ( ε ) holds true.
By setting v = u ˜ ε φ ε H 0 1 ( Ω ) as the test function in (A14), where φ ε U ˜ ε is the lifting of the Dirichlet data ε 2 g 1 on Ω , we have
Ω α ε σ ( u ˜ ε ) · s u ˜ ε + Ω ρ ε k | u ˜ ε | 2 = ε 2 Ω g 1 α n u ˜ ε E 1 + ε B ε g 2 u ˜ ε E 2 Ω ρ ε k w ε u ˜ ε E 3 .
Therefore, from the Cauchy–Schwarz inequality, there are
| E 1 | = ε 2 Ω g 1 α n u ˜ ε ε 2 g 1 L 2 ( Ω ) n u ˜ ε L 2 ( Ω ) C 1 ε 2 u ˜ ε H 1 ( Ω ) ,
| E 2 | = ε B ε g 2 u ˜ ε ε g 2 L 2 ( B ε ) u ˜ ε L 2 ( B ε ) C 2 ε 3 / 2 u ˜ ε H 1 ( Ω ) ,
| E 3 | = Ω ρ ε k w ε u ˜ ε w ε L 2 ( Ω ) u ˜ ε L 2 ( Ω ) C 3 ε 2 | ln ( ε ) | u ˜ ε H 1 ( Ω ) .
From these last results, we obtain
Ω ρ ε σ ( u ˜ ε ) · s u ˜ ε + Ω ρ ε k | u ˜ ε | 2 C 5 ( ε 2 + ε 3 / 2 + ε 2 | ln ( ε ) | ) u ˜ ε H 1 ( Ω ) .
By taking into account the coercivity of the bilinear form on the left-hand side of the above inequality, there is
c u ˜ ε H 1 ( Ω ) 2 Ω ρ ε σ ( u ˜ ε ) · s u ˜ ε + Ω ρ ε k | u ˜ ε | 2 ,
which leads to the result with constants c and C 5 independent of ε .
Corollary A1.
Let u and u ε be solutions of problems (2) and (7), respectively. Then,
u ε u L 2 ( Ω ) = o ( ε ) .
By taking into account the ansatz (A1) and the triangular inequality, it follows that
u ε u L 2 ( Ω ) = w ε + u ˜ ε L 2 ( Ω ) w ε L 2 ( Ω ) + u ˜ ε L 2 ( Ω ) w ε L 2 ( Ω ) + u ˜ ε H 1 ( Ω ) = o ( ε ) .
where we have used Lemma A1.
Before proceeding, let us subtract (2) from (7). After a simple manipulation by taking into account the contrasts (Table 2 and Table 3), one can obtain
Ω α σ ( u ε u ) · s v + Ω ρ k ( u ε u ) v = B ε ( 1 γ α ) α σ ( u ε ) · s v + B ε ( 1 γ ρ ) ρ k u ε v B ε ( 1 γ β ) β f v .

Appendix A.1. Proof of Theorem 1

By subtracting G ( u ) from G ε ( u ε ) , there is
G ε ( u ε ) G ( u ) = 2 Ω ρ k ( u ε u ) u A 1 B ε ( 1 γ ρ ) ρ k | u ε | 2 A 2 + Ω ρ k | u ε u | 2 E 1 ( ε ) ,
with the remainder E 1 ( ε ) bounded as follows:
| E 1 ( ε ) | C 1 u ε u L 2 ( Ω ) 2 = o ( ε 2 ) ,
where we have used Corollary A1. The integral A 2 can be trivially expanded as follows:
A 2 = π ε 2 ( 1 γ ρ ) ρ k | u | 2 ( x ^ ) + B ε ( 1 γ ρ ) ρ k | u ε u | 2 E 2 ( ε ) + 2 B ε ( 1 γ ρ ) ρ k ( u ε u ) u E 3 ( ε ) + B ε ( 1 γ ρ ) ρ k [ | u | 2 | u ( x ^ ) | 2 ] E 4 ( ε ) .
with remainders E 2 ( ε ) , E 3 ( ε ) , and E 4 ( ε ) bounded as follows:
| E 2 ( ε ) | C 2 u ε u L 2 ( Ω ) 2 = o ( ε 2 ) ,
| E 3 ( ε ) | C 3 ε u ε u L 2 ( Ω ) = o ( ε 2 ) ,
| E 4 ( ε ) | C 4 x x ^ L 2 ( B ε ) 2 = o ( ε 2 ) .
where we have used Corollary A1 together with the interior elliptic regularity of function u. Now, let us set v = q in (A23) and v = u ε u in the adjoint Equation (5). After comparing the obtained results, the integral A 1 can be rewritten as
A 1 = B ε ( 1 γ α ) α σ ( u ε ) · s q A 3 B ε ( 1 γ ρ ) ρ k u ε q A 4 + B ε ( 1 γ β ) β f q A 5 .
The integrals A 5 and A 6 are trivially expanded as
A 4 = π ε 2 ( 1 γ ρ ) ρ k u q ( x ^ ) + B ε ( 1 γ ρ ) ρ k ( u ε u ) q E 5 ( ε ) + B ε ( 1 γ ρ ) ρ k [ u q u q ( x ^ ) ] E 6 ( ε ) ,
A 5 = π ε 2 ( 1 γ β ) β f q ( x ^ ) + B ε ( 1 γ β ) β f [ q q ( x ^ ) ] E 7 ( ε ) .
with the remainders E 5 ( ε ) , E 6 ( ε ) , and E 7 ( ε ) bounded as follows:
| E 5 ( ε ) | ε u ε u L 2 ( Ω ) = o ( ε 2 ) ,
| E 6 ( ε ) | ε x x ^ L 2 ( B ε ) = o ( ε 2 ) ,
| E 7 ( ε ) | ε x x ^ L 2 ( B ε ) = o ( ε 2 ) ,
where we have used Corollary A1 and the interior elliptic regularity of functions u and q. The integrals A 3 and A 4 can be developed in the following way,
A 3 + A 4 = B ε ( 1 γ α ) α σ ( u ) · s q A 6 + B ε ( 1 γ α ) α σ ( w ε ) · s q A 7 + B ε ( 1 γ α ) α σ ( u ˜ ε ) · s q E 8 ( ε ) ,
where we have introduced the ansatz (A1). Therefore,
A 6 = π ε 2 ( 1 γ α ) α σ ( u ) · s q ( x ^ ) + B ε ( 1 γ α ) α [ σ ( u ) · s q σ ( u ) · s q ( x ^ ) ] E 9 ( ε ) ,
with remainders E 8 ( ε ) and E 9 ( ε ) bounded as follows:
| E 8 ( ε ) | ε u ˜ ε H 1 ( Ω ) = o ( ε 2 ) ,
| E 9 ( ε ) | ε u ˜ ε H 1 ( Ω ) = o ( ε 2 ) .
where we have used Lemma A1 together with the interior elliptic regularity of functions u and q. The last two integral A 7 can be rewritten as
A 7 = π ε 2 ( 1 γ α ) α P σ ( u ) · s q ( x ^ ) + B ε ( 1 γ α ) α σ ( w ε ) · s ( q q ( x ^ ) ) E 10 ( ε ) ,
where we have used the explicit solution (A6)–(A8). The remainder E 10 ( ε ) can be bounded as follows:
| E 10 ( ε ) | C 5 s w ε L 2 ( B ε ) x x ^ L 2 ( B ε ) = o ( ε 2 ) .
Finally, after collecting the obtained results, we have
G ε ( u ε ) G ( u ) = π ε 2 [ 2 α P σ ( u ) · s q ( x ^ ) + ( 1 γ ρ ) ρ k u ( u + q ) ( x ^ ) ( 1 γ β ) β f q ( x ^ ) ] + i = 1 10 E i ( ε ) ,
where the remainders E i ( ε ) = o ( ε 2 ) for i = 1 10 . □

Appendix A.2. Proof of Theorem 2

Let us subtract J ( u ) from J ε ( u ε ) to obtain
J ε ( u ε ) J ( u ) = 2 Ω α σ ( u ε u ) · s u B ε ( 1 γ α ) α σ ( u ε ) · s u ε + Ω α σ ( u ε u ) · s ( u ε u ) B 1 .
By setting v = u ε u as the test function in (A23), the integral B 1 can be rewritten, after some manipulations, as
B 1 = B ε ( 1 γ α ) α σ ( u ε ) · s ( u ε u ) + E 11 ( ε ) .
The remainder E 11 ( ε ) is defined as
E 11 ( ε ) = B ε ( 1 γ ρ ) ρ k | u ε u | 2 + B ε ( 1 γ ρ ) ρ k u ( u ε u ) B ε ( 1 γ β ) β f ( u ε u ) Ω ρ k | u ε u | 2 ,
which can be bounded as follows:
| E 11 ( ε ) | C 1 ( ε + u ε u L 2 ( B ε ) + s ( u ε u ) L 2 ( B ε ) ) u ε u L 2 ( B ε )   + C 2 ( u ε u L 2 ( Ω ) + s ( u ε u ) L 2 ( Ω ) ) u ε u L 2 ( Ω )   C 3 u ε u L 2 ( Ω ) u ε u H 1 ( Ω ) = o ( ε 2 ) ,
where we have used the Cauchy–Schwarz inequality together with Lemma 1 and Corollary A1. Therefore, Equation (A43) becomes
J ε ( u ε ) J ( u ) = 2 Ω α σ ( u ε u ) · s u B 2 B ε ( 1 γ α ) α σ ( u ε ) · s u B 3 + E 11 ( ε ) .
From the ansatz (A1), integral B 3 can be written as
B 3 = B ε ( 1 γ α ) α σ ( u ) · s u B 4 + B ε ( 1 γ α ) α σ ( w ε ) · s u B 5 + B ε ( 1 γ α ) α σ ( u ) · s u ˜ ε E 12 ( ε ) ,
with the remainder E 12 ( ε ) bounded as follows:
| E 12 ( ε ) | C 1 ε s u ˜ ε L 2 ( B ε ) C 2 ε u ˜ ε H 1 ( Ω ) = o ( ε 2 )
where we have used Lemma A1. The integrals B 4 and B 5 can be trivially expanded as follows:
B 4 = π ε 2 ( 1 γ α ) α σ ( u ( x ^ ) ) · s u ( x ^ ) + B ε ( 1 γ α ) α ( σ ( u ) · s u σ ( u ( x ^ ) ) · s u ( x ^ ) E 13 ( ε ) ,
B 5 = π ε 2 ( 1 γ α ) α P σ ( u ) · s u ( x ^ ) + B ε ( 1 γ α ) α σ ( w ε ) · ( s u s u ( x ^ ) ) E 14 ( ε ) ,
where we have used the explicit solution (A6)–(A8). The remainders E 13 ( ε ) and E 14 ( ε ) can be bounded as follows:
| E 13 ( ε ) | C 1 ε x x ^ L 2 ( B ε ) = o ( ε 2 ) ,
| E 14 ( ε ) | C 2 σ ( w ε ) L 2 ( B ε ) x x ^ L 2 ( B ε ) = o ( ε 2 ) .
Now, let us set v = p in (A23) and v = u ε u in the adjoint Equation (6). After comparing the obtained results, the integral B 2 can be rewritten as
B 2 = B ε ( 1 γ α ) α σ ( u ε ) · s p B 6 B ε ( 1 γ ρ ) ρ k u ε p B 7 + B ε ( 1 γ β ) β f p B 8 .
The integrals B 7 and B 8 are trivially expanded as
B 7 = π ε 2 ( 1 γ ρ ) ρ k u p ( x ^ ) + B ε ( 1 γ ρ ) ρ k ( u ε u ) p E 15 ( ε ) + B ε ( 1 γ ρ ) ρ k [ u p u p ( x ^ ) ] E 16 ( ε ) ,
B 8 = π ε 2 ( 1 γ β ) β f p ( x ^ ) + B ε ( 1 γ β ) β f [ p p ( x ^ ) ] E 17 ( ε ) .
with remainders E 15 ( ε ) , E 16 ( ε ) , and E 17 ( ε ) bounded as follows:
| E 15 ( ε ) | ε u ε u L 2 ( Ω ) = o ( ε 2 ) ,
| E 16 ( ε ) | ε x x ^ L 2 ( B ε ) = o ( ε 2 ) ,
| E 17 ( ε ) | ε x x ^ L 2 ( B ε ) = o ( ε 2 ) ,
where we have used Corollary A1 and the interior elliptic regularity of functions u and p. The integrals B 6 and B 7 can be developed in the following way,
B 6 + B 7 = B ε ( 1 γ α ) α σ ( u ) · s p B 9 + B ε ( 1 γ α ) α σ ( w ε ) · s p B 10 + B ε ( 1 γ α ) α σ ( u ˜ ε ) · s p E 18 ( ε ) ,
where we have introduced the ansatz (A1). Therefore,
B 9 = π ε 2 ( 1 γ α ) α σ ( u ) · s p ( x ^ ) + B ε ( 1 γ α ) α [ σ ( u ) · s p σ ( u ) · s p ( x ^ ) ] E 19 ( ε ) .
with remainders E 18 ( ε ) and E 19 ( ε ) , bounded as follows:
| E 18 ( ε ) | ε u ˜ ε H 1 ( Ω ) = o ( ε 2 ) ,
| E 19 ( ε ) | ε u ˜ ε H 1 ( Ω ) = o ( ε 2 ) ,
where we have used Lemma A1 together with the interior elliptic regularity of functions u and p. The last integral B 11 can be rewritten as
B 11 = π ε 2 ( 1 γ α ) α P σ ( u ) · s p ( x ^ ) + B ε ( 1 γ α ) α σ ( w ε ) · s ( p p ( x ^ ) ) E 20 ( ε ) ,
where we have used the explicit solution (A6)–(A8). The remainders E 20 ( ε ) can be bounded as follows:
| E 20 ( ε ) | C 1 σ ( w ε ) L 2 ( B ε ) x x ^ L 2 ( B ε ) = o ( ε 2 ) .
Finally, after collecting the obtained results, we have
J ε ( u ε ) J ( u ) = π ε 2 [ 2 α P σ ( u ) · s ( u + p ) ( x ^ ) + ( 1 γ ρ ) ρ k u p ( x ^ ) ( 1 γ β ) β f p ( x ^ ) ] + i = 11 20 E i ( ε ) ,
where the remainders E i ( ε ) = o ( ε 2 ) , for i = 11 20 . □

References

  1. Masur, E.F.; Mroz, Z. Non-stationary optimality conditions in structural design. Int. J. Solids Struct. 1979, 15, 503–512. [Google Scholar] [CrossRef]
  2. Haug, E.J.; Rousselet, B. Design sensitivity analysis in structural mechanics. II. eigenvalue variationsi. J. Struct. Mech. 1979, 8, 161–186. [Google Scholar] [CrossRef]
  3. Seyranian, A.P.; Lund, E.; Olhoff, N. Multiple eigenvalues in structural optimization problems. Struct. Optim. 1994, 8, 207–227. [Google Scholar] [CrossRef]
  4. Jensen, J.S.; Pedersen, N.L. On maximal eigenfrequency separation in two-material structures: The 1D and 2D scalar cases. J. Sound Vib. 2006, 289, 967–986. [Google Scholar] [CrossRef]
  5. Gravesen, J.; Evgrafov, A.; Nguyen, D.M. On the sensitivities of multiple eigenvalues. Struct Multidiscip Optim 2011, 44, 583–587. [Google Scholar] [CrossRef]
  6. Torii, A.; Faria, J.R. Structural optimization considering smallest magnitude eigenvalues: A smooth approximation. J. Braz. Soc. Mech. Sci. Eng. 2017, 39, 1745–1754. [Google Scholar] [CrossRef]
  7. Ammari, H.; Khelifi, A. Electromagnetic scattering by small dielectric inhomogeneities. J. Math. Pures Appl. 2003, 82, 749–842. [Google Scholar] [CrossRef]
  8. Nazarov, S.A.; Sokołowski, J. Spectral problems in the shape optimisation. Singular boundary perturbations. Asymptot. Anal. 2008, 56, 159–204. [Google Scholar]
  9. Amstutz, S. Augmented Lagrangian for cone constrained topology optimization. Comput. Optim. Appl. 2011, 101–122. [Google Scholar] [CrossRef]
  10. Carvalho, F.; Ruscheinsky, D.; Giusti, S.; Anflor, C.; Novotny, A. Topological Derivative-Based Topology Optimization of Plate Structures Under Bending Effects. Struct. Multidiscip. Optim. 2020, 63, 617–630. [Google Scholar] [CrossRef]
  11. Ruscheinsky, D.; Carvalho, F.S.; Anflor, C.; Novotny, A.A. Topological asymptotic analysis of a diffusive–convective–reactive problem. Eng. Comput. 2020, 38, 477–500. [Google Scholar] [CrossRef]
  12. Takezawa, A.; Nishiwaki, S.; Kitamura, M. Shape and topology optimization based on the phase field method and sensitivity analysis. J. Comput. Phys. 2010, 229, 2697–2718. [Google Scholar] [CrossRef]
  13. Qian, M.; Hu, X.; Zhu, S. A phase field method based on multi-level correction for eigenvalue topology optimization. Comput. Methods Appl. Mech. Eng. 2022, 401, 115646. [Google Scholar] [CrossRef]
  14. Amstutz, S.; Andrä, H. A new algorithm for topology optimization using a level-set method. J. Comput. Phys. 2006, 216, 573–588. [Google Scholar] [CrossRef]
  15. Novotny, A.A.; Sokołowski, J. Topological Derivatives in Shape Optimization; Interaction of Mechanics and Mathematics; Springer: Berlin/Heidelberg, Germany, 2013; p. 324. [Google Scholar]
  16. Novotny, A.A.; Sokolowski, J. An Introdution to the Topological Derivative Method; Springer Briefs in Mathematics; Springer: New York, NY, USA, 2020. [Google Scholar]
Figure 1. Case A: initial domain (a) and displacement (b).
Figure 1. Case A: initial domain (a) and displacement (b).
Mathematics 12 02762 g001
Figure 2. Case B: initial domain (a) and displacement (b).
Figure 2. Case B: initial domain (a) and displacement (b).
Mathematics 12 02762 g002
Figure 3. Case C: initial domain (a) and displacement (b).
Figure 3. Case C: initial domain (a) and displacement (b).
Mathematics 12 02762 g003
Figure 4. Optimized topology for Case A.
Figure 4. Optimized topology for Case A.
Mathematics 12 02762 g004
Figure 5. Optimized topology.
Figure 5. Optimized topology.
Mathematics 12 02762 g005
Figure 6. Optimized topology for Case B.
Figure 6. Optimized topology for Case B.
Mathematics 12 02762 g006
Figure 7. Optimized topology for Case C.
Figure 7. Optimized topology for Case C.
Mathematics 12 02762 g007
Figure 8. Normalized first eigenvalue λ 1 / λ 1 0 history.
Figure 8. Normalized first eigenvalue λ 1 / λ 1 0 history.
Mathematics 12 02762 g008
Figure 9. Normalized second eigenvalue λ 1 / λ 2 history.
Figure 9. Normalized second eigenvalue λ 1 / λ 2 history.
Mathematics 12 02762 g009
Figure 10. Shape function history.
Figure 10. Shape function history.
Mathematics 12 02762 g010
Figure 11. Volume fraction history.
Figure 11. Volume fraction history.
Mathematics 12 02762 g011
Figure 12. Cases D1, D2, D3, and D4: initial domain (a) and displacement (b).
Figure 12. Cases D1, D2, D3, and D4: initial domain (a) and displacement (b).
Mathematics 12 02762 g012
Figure 13. Comparison of designs obtained for Cases D1, D2, D3, and D4.
Figure 13. Comparison of designs obtained for Cases D1, D2, D3, and D4.
Mathematics 12 02762 g013
Figure 14. Cases D1, D2, D3, and D4: shape function history.
Figure 14. Cases D1, D2, D3, and D4: shape function history.
Mathematics 12 02762 g014
Figure 15. Cases D1, D2, D3, and D4: volume fraction history.
Figure 15. Cases D1, D2, D3, and D4: volume fraction history.
Mathematics 12 02762 g015
Figure 16. Cases D1, D2, D3, and D4: normalized first eigenvalue λ 1 / λ 1 0 history.
Figure 16. Cases D1, D2, D3, and D4: normalized first eigenvalue λ 1 / λ 1 0 history.
Mathematics 12 02762 g016
Figure 17. Cases D1, D2, D3, and D4: normalized second eigenvalue λ 1 / λ 2 history.
Figure 17. Cases D1, D2, D3, and D4: normalized second eigenvalue λ 1 / λ 2 history.
Mathematics 12 02762 g017
Table 1. Values of α , ρ , and β .
Table 1. Values of α , ρ , and β .
α ρ β
Ω ω α 0 ρ 0 β 0
ω α 1 ρ 1 β 1
Table 2. Values of α ε , ρ ε , and β ε .
Table 2. Values of α ε , ρ ε , and β ε .
α ε ρ ε β ε
Ω B ε α ρ β
B ε γ α α γ ρ ρ γ β β
Table 3. Values of γ α , γ ρ , and γ β .
Table 3. Values of γ α , γ ρ , and γ β .
γ α γ ρ γ β
Ω ω α 1 / α 0 ρ 1 / ρ 0 β 1 / β 0
ω α 0 / α 1 ρ 0 / ρ 1 β 0 / β 1
Table 4. Considered cases.
Table 4. Considered cases.
CasesGridElementsNodes ν μ
A 30 × 90 10,80055210.251.4
B 30 × 180 21,60011,0110.30.7
C 10 × 40 16008510.31.5
Table 5. Penalty parameter.
Table 5. Penalty parameter.
Case D1Case D2Case D3Case D4
μ 0.5 1.0 1.5 2.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Carvalho, F.S.; Anflor, C.T.M. The Concept of Topological Derivative for Eigenvalue Optimization Problem for Plane Structures. Mathematics 2024, 12, 2762. https://doi.org/10.3390/math12172762

AMA Style

Carvalho FS, Anflor CTM. The Concept of Topological Derivative for Eigenvalue Optimization Problem for Plane Structures. Mathematics. 2024; 12(17):2762. https://doi.org/10.3390/math12172762

Chicago/Turabian Style

Carvalho, Fernando Soares, and Carla Tatiana Mota Anflor. 2024. "The Concept of Topological Derivative for Eigenvalue Optimization Problem for Plane Structures" Mathematics 12, no. 17: 2762. https://doi.org/10.3390/math12172762

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop