Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Electromagnetic-acoustic biphysical cloak designed through topology optimization

Open Access Open Access

Abstract

Various strategies have been proposed to achieve invisibility cloaking, but usually only one phenomenon is controlled by each device. Cloaking an object from two different waves, such as electromagnetic and acoustic waves, is a challenging problem, if not impossible, to be achieved using transformation theory and metamaterials, which are the major approaches in physics. Here, by developing topology optimization for controlling both electromagnetic and acoustic waves, we present a multidisciplinary attempt for designing biphysical cloaks with triple-wave cloaking capabilities, specifically for Ez- and Hz-polarized waves and acoustic wave. The topology-optimized biphysical cloak cancels the scattering of the three waves and reproduces the original propagating waves as if nothing is present, thus instilling the desired cloaking capability. In addition, we describe cloaking structures for multiple incident directions of the three waves and structures that work for both electromagnetic waves and sound waves of different wavelengths.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Cloaking an object from two different physical flows has been a focus since the pioneering work to develop a biphysical cloak manipulating both heat flux and direct current (DC) was proposed [1]. The theoretical concept behind biphysical cloaking in diffusive systems was, after a while, experimentally demonstrated using bifunctional metamaterials [2], bilayer shell systems [3] and a combination of passive and active schemes [4]. These technologies have been extended to invisible sensing [5] and independent thermal and electrical fields manipulations [6] in multiple-physical diffusive fields. Recently, computational inverse design based on topology optimization was demonstrated for thermal-electrical biphysical cloaks [7]. The two diffusive systems are widely known to exhibit very similar phenomena because both thermal and DC electrical conduction satisfy Laplace’s equation in steady states, and insulators of both are modeled using Neumann boundary conditions.

While development of biphysical cloaking in diffusive systems has been successful, cloaking an object from two different wave types is much more difficult. The performances of cloaking devices based on metamaterials [8,9] for a wave system are far from ideal because transformation optics-based methods [1012] require unrealistic physical parameters and inhomogeneous media with complex material profiles. Constructing an inhomogeneous anisotropic device of this type remains very difficult, offering little advantage for practical applications. The earliest attempt for biphysical cloaking in wave systems proposed a structured cloak made of aluminum [13]; shortly thereafter, a carpet cloak based on a metasurface controlling local reflection phases was developed [14]. Recently, multi-physics surface transformations [15] were formulated offering simultaneous manipulations of polarized electromagnetic and acoustic waves. These three theories are based on aluminum structures used in constructing biphysical cloaks; as a perfect electric conductor (PEC) to $H_z$-polarized waves and hard wall to acoustic waves in air, both are governed by the Helmholtz equation applying Neumann boundary conditions. Therefore, the three theories can not be applied when cloaking an object from $E_z$-polarized wave, for which Dirichlet boundary conditions need to be imposed on the PEC. Additionally, the PEC is not suitable for modeling materials with low electrical conductivity, and very few materials (e.g., metals) can thus be used to construct the cloak. To achieve an ideal biphysical cloak, cloaking an object is essential for not only $H_z$-polarized and acoustic waves but also $E_z$-polarized waves. That is, triple-wave cloaking is required but is such cloaking possible for electromagnetic and acoustic waves?

Using transformation theory and metamaterials, solving this problem is very challenging, if not impossible. In recent years, remarkable progress has been made in informatics, which involves data analysis, optimization, and machine learning, and is seen as a powerful tool for overcoming such challenging problems. These innovative algorithms are not limited to information science and are being applied to the design of wave-controlling devices in the fields of optics [1618] and electromagnetism. Topology optimization, the most flexible structural optimization in allowing creation of pores, has performed particularly well in designing high-performance metastructures for cloaks in not only diffusive systems [19,20] but also single-propagating-wave systems for both electromagnetic cloaks [2125] and acoustic cloaks [2629]. However, existing topology optimization approaches produce only metastructures designed to cloak objects for either electromagnetic waves or acoustic waves.

In this work, we propose a topology optimization for biphysical cloaks with triple-wave cloaking capabilities, specifically for $E_z$- and $H_z$-polarized waves and acoustic wave. For this purpose, we set up three objective functions, one for each of the three waves. For polarization-independent electromagnetic cloaking, structures of optimal topology made of acrylonitrile-butadiene-styrene (ABS) resin that surrounded a PEC obstacle were explored by solving the Helmholtz equations; Dirichlet and Neumann boundary conditions were imposed on the propagating $E_z$- and $H_z$-polarized waves, respectively. In evaluating acoustic cloaking, the acoustic-elastic coupled problem were solved taking the elasticity into account. To optimize the structure as a biphysical cloak, the covariance matrix adaptation evolution strategy (CMA-ES) [30] is implemented; using level-set methods, the aim is to suppress scattering of all $E_z$- and $H_z$-polarized waves and the acoustic wave.

2. Formulation

2.1 Optimization scheme for triple wave cloaking

To attain biphysical cloaking in the wave system [Fig. 1(a)], topology optimization involves suppressing both electromagnetic and acoustic scattering produced by an obstacle $\Omega _\mathrm {obs}$, here made of aluminum; the fixed design domain $\Omega _\mathrm {D}$, in which the ABS structure $\Omega _\mathrm {s}$ is transformed to affect biphysical cloaking, is set to cover $\Omega _\mathrm {obs}$. Structural symmetries are imposed on $x$ and $y$ axes. To establish wave scattering in the open region, we configure a perfectly matched layer (PML), $\Omega _\mathrm {PML}$, imposing on it the absorbing boundary condition. The interfaces between free space, $\Omega _\mathrm {D}\backslash \Omega _\mathrm {s}$, and the structure, $\Omega _\mathrm {s}$, are expressed as iso-surfaces of a level-set function [Fig. 1(b)], from which finite elements are generated. We note that, by using a level set method, no grayscale is included in the structure, resulting in a fully binarized and optimized structure with crisp structural boundaries; hence, there is no need to use a filtering scheme to remove a grayscale.

Table 1 lists the material parameters used in our topology optimization. The relative permittivity of the ABS is set assuming the cloaking of an object from microwaves.

 figure: Fig. 1.

Fig. 1. Schematic of a topology optimization for an electromagnetic-acoustic biphysical cloak. Domain sizes are $R_\mathrm {obs}=R_\mathrm {D}/3$, $L_x=R_\mathrm {D}\times 2$, $L_y=R_\mathrm {D}\times 3/2$, and $L_\mathrm {grid}=R_\mathrm {D}/70$.

Download Full Size | PDF

Tables Icon

Table 1. Material Properties in the Optimization.

2.2 Objective functions

The materials occupying $\Omega _\mathrm {s}$ that constitute the cloak are both dielectric and elastic and minimize the scattering of electromagnetic and acoustic waves in the outer domain $\Omega _\mathrm {out}$. The objective function determining the design of the electromagnetic-acoustic biphysical cloak is defined as

$$\Psi_X= \frac{1}{\Psi_X^\mathrm{n}} \int_\mathrm{\Omega_\mathrm{out}} \big|X^\mathrm{s}\big|^2 d\Omega,$$
where $X\in (E, H, p)$ signifies the physical field, specifically, $E$ for the electric field of the $E_z$-polarized wave, $H$ for the magnetic field of the $H_z$-polarized wave, and $p$ for the sound pressure of the acoustic wave. The scattered wave of the field is denoted $X^\mathrm {s}\in \{E^\mathrm {s}, H^\mathrm {s}, p^\mathrm {s}\}$. The total wave is given as $X=X^\mathrm {s}+X^\mathrm {i}$ with $X^\mathrm {i}\in \{E^\mathrm {i}, H^\mathrm {i}, p^\mathrm {i}\}$ denoting the incident wave. Then, canceling the scattered field $X^\mathrm {s}$ makes the total field $X$ close to the incident field $X^\mathrm {i}$ and reproduces $X^\mathrm {i}$ as if nothing exists, thus realizing cloaking. With subscript $X$, $\Psi _X$ denotes the corresponding objective function, its normalization $\Psi _X^\mathrm {n}$ being
$$\Psi_X^\mathrm{n}=\int_\mathrm{\Omega_\mathrm{out}} \big| X_\mathrm{bare}^\mathrm{s} \big|^2 d\Omega,$$
where $X_\mathrm {bare}^\mathrm {s}\in (E_\mathrm {bare}^\mathrm {s}, H_\mathrm {bare}^\mathrm {s}, p_\mathrm {bare}^\mathrm {s})$ signifies the scattered wave from the bare cylinder without the cloak.

2.3 Governing equations and boundary conditions

Although the objective functions that determine the canceling of the scattered $E_z$- and $H_z$-polarized waves and the acoustic wave may be expressed in the same form (1), the electromagnetic waves and acoustic wave satisfy different governing equations and boundary conditions. The $E_z$- and $H_z$-polarized waves satisfy the following Helmholtz equations;

$$E_{,ii}+\omega_\mathrm{e}^2\epsilon c^{{-}2}E=0 \hspace{7mm}\mathrm{in}\ \Omega_\mathrm{D}\backslash \Omega_\mathrm{s}, \Omega_\mathrm{s}, \Omega_\mathrm{out}, $$
$$ \epsilon^{{-}1}H_{,ii}+\omega_\mathrm{e}^2 c^{{-}2}H=0 \hspace{7mm}\mathrm{in}\ \Omega_\mathrm{D}\backslash \Omega_\mathrm{s}, \Omega_\mathrm{s}, \Omega_\mathrm{out}, $$
where $\epsilon$ denotes the relative permittivity, $\omega _\mathrm {e}$ the angular frequency of the electromagnetic wave, and $c$ the speed of the electromagnetic wave in air; here, Einstein’s summation convention is used. We assume the obstacle $\Omega _\mathrm {obs}$ to be a PEC. Hence, the PEC boundary condition is implemented for $E_z$ and $H_z$ polarizations by setting
$$ E=0 \hspace{7mm}\mathrm{on}\ \Gamma_\mathrm{obs}^\mathrm{s}, \Gamma_\mathrm{D}^\mathrm{obs}, $$
$$ n^\mathrm{PEC}_i H_{,i}=0 \hspace{7mm}\mathrm{on}\ \Gamma_\mathrm{obs}^\mathrm{s}, \Gamma_\mathrm{D}^\mathrm{obs}, $$
where $n^\mathrm {PEC}_i$ is the unit normal to the surface of the PEC [Fig. 1(c,d)]. On the boundaries between the ABS and air, the tangential component of the electromagnetic field is conserved; i.e. the continuity of $E$ for the $E_z$ polarization and that of $H$ for the $H_z$ polarization are satisfied, respectively, as the boundary condition imposed on $\Gamma _\mathrm {D}^\mathrm {s}$ in the electromagnetic analyses of the presented two-dimensional system.

The governing equations for sound pressure $p$ in an acoustic medium, and the stress $\sigma _{ij}$ and displacements $u_{i}$ in the elastic media, are

$$\rho_\mathrm{a}^{{-}1}p_{,ii}+\omega_\mathrm{a}^2\kappa_\mathrm{a}^{{-}1}p=0 \hspace{7mm}\mathrm{in}\ \Omega_\mathrm{D}\backslash \Omega_\mathrm{s}, \Omega_\mathrm{out}, $$
$$ \sigma_{ji,j}+\rho_\mathrm{e}\omega_\mathrm{a}^2u_i=0 \hspace{7mm}\mathrm{in}\ \Omega_\mathrm{s}, \Omega_\mathrm{obs}, $$
where $\rho _\mathrm {a}$ and $\rho _\mathrm {e}$ denote the densities of the acoustic and elastic media, $\omega _\mathrm {a}$ the angular frequency of the acoustic wave, and $\kappa _\mathrm {a}$ the bulk modulus of the acoustic medium. At the interface between the two media, the boundary conditions are
$$p_{,i}n_i^\mathrm{s}=\omega_\mathrm{a}^2\rho_\mathrm{a}u_in_i^\mathrm{s} \hspace{7mm}\mathrm{on}\ \Gamma_\mathrm{D}^\mathrm{s}, \Gamma_\mathrm{D}^\mathrm{obs}, $$
$$ \sigma_{ji}(\boldsymbol{u})n_i^\mathrm{s}={-}pn_i^\mathrm{s} \hspace{13mm}\mathrm{on}\ \Gamma_\mathrm{D}^\mathrm{s}, \Gamma_\mathrm{D}^\mathrm{obs}, $$
where $n_i^\mathrm {s}$ represents outward unit normal to the boundaries of the elastic domain [Fig. 1(b)] and $\boldsymbol {u}$ the displacement vector.

2.4 Level set method and regularized fitness

The free space and structural domains, $\Omega _\mathrm {D}\backslash \Omega _\mathrm {s}$ and $\Omega _\mathrm {s}$, are outlined using a level-set function $\phi (\boldsymbol {x})$ defined as

$$\begin{aligned} \Omega_\mathrm{D}\backslash \Omega_\mathrm{s} & = \{ \boldsymbol{x}\ |\ -1 \leq \phi(\boldsymbol {x}) < 0 \}, \\ \Gamma_\mathrm{D}^\mathrm{s} & = \{ \boldsymbol{x}\ |\ \phi(\boldsymbol {x}) = 0 \}, \\ \Omega_\mathrm{s}\backslash \Gamma_\mathrm{D}^\mathrm{s} & = \{ \boldsymbol{x}\ |\ 0 < \phi(\boldsymbol {x}) \leq 1 \}. \end{aligned} $$

Here, the interfaces between the free space and structural domains, $\Gamma _\mathrm {D}^\mathrm {s}$, are expressible as iso-surfaces of the level-set function [Fig. 1(b)]. The optimal $\boldsymbol {\phi }=\{ \phi _1, \ldots, \phi _j, \ldots, \phi _n \}$, the set of discretized level-set functions, is explored using CMA-ES [30] (samplings $\lambda =400$) with box-constraint handling [31], $-1\leq \phi _j \leq 1$, for infimizing regularized fitness,

$$\underset{\phi}{\mathrm{inf}}\qquad F= \max\big( \Psi_E, \Psi_H, \Psi_p, \tau L \big),$$
where $L$ denotes the normalized perimeter length of the structure and $\tau$ the coefficient of regularization, specifically, the ratio between the maximum of the three objective functions and $L$.

2.5 CMA-ES

CMA-ES [30,32] is one of the most powerful evolutionary computation approaches for searching optimal design variables $\boldsymbol{\phi }$, and the algorithm of the CMA-ES is based on the adaptation of the distribution parameters for normally distributed sampling (i.e., the shape, size, and center of the distribution) to the landscape of the minimized fitness $F$. The CMA-ES is known to be robust against optimization problems having difficult properties [33], such as multimodality and the interdependence of design variables. The CMA-ES enables designers to realize well-performing configurations [3436] in topology optimizations with the use of box constraint handling [31,37] without implementing a trial-and-error approach to make initial guesses. The default values of strategy parameters of CMA-ES have been well studied [38] and the parameters do not need to be adjusted. The only hyper-parameter that is advised to tune is the population size, $\lambda$, which is the number of candidate solutions. It is empirically known that a greater population size tends to be more robust against local minima [39]. Because we can parallelize the fitness evaluation of $\lambda$ candidate solutions and the number of generations of the algorithm reduces as we increase $\lambda$ [40], we can accelerate the optimization wall clock time through parallel implementation. We used the following CMA-ES algorithm.

The distribution parameters are initialized at the first generation $g=0$ as

$$\boldsymbol{m}^{(0)} =0.5(b_\mathrm{u} +b_\mathrm{l}) \mathbf{1}_n, \hspace{3mm} \sigma^{(0)}=0.3(b_\mathrm{u}-b_\mathrm{l}), \hspace{3mm} \boldsymbol{C}^{(0)} =\boldsymbol{I}_n, \hspace{3mm} \boldsymbol{p}_{\mathrm \sigma}^{(0)}= \boldsymbol{0}, \hspace{3mm} \boldsymbol{p}_{\boldsymbol C}^{(0)}=\boldsymbol{0}, \hspace{3mm}$$
where $\boldsymbol{m}^{(g)}$ is the mean vector of the sampling distribution, $b_\mathrm {u}$ is the upper bound of the design variable, $b_\mathrm {l}$ is the lower bound of the design variable, $\mathbf {1}_n$ is an $n$-dimensional vector with all elements being $1$, $\sigma ^{(g)}$ is the distribution spread of the sampling, $\boldsymbol{C}^{(g)}$ is the covariance matrix of the sampling distribution, ${\boldsymbol I}_n$ is the identity matrix of size $n$, and $\boldsymbol{p}_{\mathrm \sigma }^{(g)}$ and $\boldsymbol{p}_{\boldsymbol C}^{(g)}$ are so-called evolution paths. The candidate solutions for sampling are generated from the distribution parameters as
$$\boldsymbol{\phi}_i^{(g+1)}=\boldsymbol{m}^{(g)} + \sigma^{(g)}\boldsymbol{BD} \boldsymbol{z}_{i}^{(g+1)},$$
where $\boldsymbol{z}_{i}^{(g+1)}$ is a random vector generated from a multivariate normal distribution, denoted as $\boldsymbol{z}_i^{(g+1)} \sim \mathcal {N}(\boldsymbol{0}, \boldsymbol{I}_n)$, $\boldsymbol{B}$ is a matrix composed of the eigenvectors of $\boldsymbol{C}^{(g)}$, and $\boldsymbol{D}$ is a diagonal matrix in which the diagonal elements equal the square root of the eigenvalues of $\boldsymbol{C}^{(g)}$. The matrices $\boldsymbol{B}$ and $\boldsymbol{D}$ are obtained through the eigen-decomposition of $\boldsymbol{C}^{(g)}$ as
$$\boldsymbol{C}^{(g)} = \boldsymbol{B}\boldsymbol{D}^2\boldsymbol{B}^\top\enspace.$$

The candidate solutions are ranked using $F$ with an adaptive penalty function for handling the box constraint [31], $-1\leq \phi _j \leq 1$. The differences between the mean vector and $i$-th-ranked solution denoted $\boldsymbol{\phi }_{i:\lambda }^{(g+1)}$ are scaled as

$$\boldsymbol{y}_{i:\lambda}^{(g+1)}=\left(\boldsymbol{\phi}_{i:\lambda}^{(g+1)}-\boldsymbol{m}^{(g)}\right)\Big/\sigma^{(g)}.$$

The distribution parameters are updated following the scheme

$$\boldsymbol{m}^{(g+1)} = \boldsymbol{m}^{(g)} + \sigma^{(g)} \boldsymbol{m}_{\boldsymbol y}, $$
$$\sigma^{(g+1)} = \sigma^{ (g)}\ \exp\Biggl(\frac{c_{\mathrm \sigma}}{d_\sigma} \biggl(\frac{||\boldsymbol{p}_{\mathrm \sigma}^{(g+1)}||} {\mathbb{E}[|| \mathcal{N}(\boldsymbol{0},{\boldsymbol I}_n)||]}-1\biggr)\Biggr), $$
$$ \boldsymbol{C}^{(g+1)} =\boldsymbol{C}^{(g)}+h_\sigma^{(g+1)} c_1\bigg( \boldsymbol{p}_{\boldsymbol C}^{(g+1)} \Big(\boldsymbol{p}_{\boldsymbol C}^{(g+1)} \Big)^\top \hspace{-1mm} -\boldsymbol{C}^{(g)} \bigg) +c_\mu\hspace{-1mm}\sum_{i=1}^{\lambda}w_i \bigg( \boldsymbol{y}_{i:\lambda}^{(g+1)} \Big(\boldsymbol{y}_{i:\lambda}^{(g+1)} \Big)^\top \hspace{-1mm} -\boldsymbol{C}^{(g)}\bigg), $$
where $c_1$ and $c_\mu$ represent the learning rates of rank-one and rank-$\mu$ updates, respectively, $d_\sigma$ is the parameter for damping step-size adaptation, $w_i$ denotes the weight assigned to the $i$th best candidate solution, $\boldsymbol{p}_{\boldsymbol C}^{(g)}$ and $\boldsymbol{p}_{\mathrm \sigma }^{(g)}$ are evolution paths that are adapted according to
$$\boldsymbol{p}_{\boldsymbol C}^{(g+1)} = (1 - c_{\boldsymbol C})\boldsymbol{p}_{\boldsymbol C}^{(g)} +h_\sigma^{(g+1)}\sqrt{c_{\boldsymbol C}(2-c_{\boldsymbol C})\mu_\mathrm{eff}}\boldsymbol{m}_{\boldsymbol y}, $$
$$\boldsymbol{p}_{\mathrm \sigma}^{(g+1)}=(1-c_{\mathrm \sigma})\boldsymbol{p}_{\mathrm \sigma}^{(g)} + \sqrt{c_{\mathrm \sigma}(2-c_{\mathrm \sigma})\mu_\mathrm{eff}} \sqrt{\boldsymbol{C}^{(g)}}^{{-}1} \boldsymbol{m}_{\boldsymbol y}, $$
where $c_{\boldsymbol{C}}$ and $c_\sigma$ represent the learning rates for the cumulations of the rank-one update and the step size, respectively, with the notation $\sqrt {\boldsymbol{C}^{(g)}}^{-1}$ expressing
$$\sqrt{\boldsymbol{C}^{(g)}}^{{-}1}=\boldsymbol{B}\boldsymbol{D}^{{-}1}\boldsymbol{B}^\top.$$

The weighted average of $\boldsymbol{y}_{i:\lambda }^{(g+1)}$, $\boldsymbol{m}_{\boldsymbol y}$, is expressed as

$$\boldsymbol{m}_{\boldsymbol y} = \sum_{i=1}^{\mu}w_i \boldsymbol{y}_{i:\lambda}^{(g+1)},$$
and $h_\sigma ^{(g+1)}$ is the Heaviside step function defined as
$$\begin{aligned}h_\sigma^{(g+1)} = \begin{cases} 1 & \ || \boldsymbol{p}_{\mathrm \sigma}^{(g+1)} || < \left(1.5 + \frac{1}{n-0.5}\right)\mathbb{E}[|| \mathcal{N}(\boldsymbol{0},{\boldsymbol I}_n)||], \\ 0 & \ \rm{otherwise}, \end{cases} \end{aligned}$$
where $\mathbb {E}[|| \mathcal {N} (\boldsymbol{0},{\boldsymbol I}_n)||]$ represents the expectation of the norm of an $n$-variate normal distribution. In fact, the expectation is approximately calculated as $\mathbb {E}[|| \mathcal {N} (\boldsymbol{0},{\boldsymbol I}_n)||]=\sqrt {n} \left (1 - \frac {1}{4n} + \frac {1}{ 21n^2}\right )$. See Refs. [31,38,41] for further details.

The convergence of the optimization is assessed according to the error of the mean vector $\boldsymbol{m}^{(g)}$, defined as

$$\Delta\boldsymbol{m}_\mathrm{error}=\frac{||\boldsymbol{m}^{(g)}-\boldsymbol{m}^{(g-1)}||}{||\boldsymbol{m}^{(g)}||} \leq 0.01.$$

The optimization process continues until the above convergence criteria are satisfied.

The internal time complexity of the CMA-ES per generation is $O(\lambda n^2)$, where $n$ is the number of design variables; i.e., the dimension of design space. The number of the fitness evaluations required is usually $O(n)$. Therefore, roughly speaking, the computational cost for the optimization algorithm itself scales as $O(n^3)$ as the dimension increases. See [40] for the scalability analysis of the CMA-ES.

3. Results

3.1 Wave scattering by a bare cylinder

A bare cylinder made of aluminum [Fig. 2(a)] is modeled as a PEC; both $E_z$- and $H_z$-polarized waves are intensively scattered from the cylinder’s surface [Fig. 2(b,c)]. The difference in interference patterns between $E_z$-polarized wave [Fig. 2(b)] and $H_z$-polarized wave [Fig. 2(c)] arise from differences in the PEC boundary conditions in Eqs. (5) and (6). Although acoustic scattering [Fig. 2(d)] is modeled as an acoustic-elastic coupled system with elasticity corresponding to aluminum, the scattering of the $H_z$-polarized wave by the PEC with Neumann boundary conditions corresponds well to that of an acoustic wave, thus supporting this model’s suitability as a hard wall.

 figure: Fig. 2.

Fig. 2. (a) Bare cylinder without a cloak and its scattering of (b) $E_z$-polarized wave, (c) $H_z$-polarized wave, and (d) acoustic wave around the cylinder; (d) mean stress $\Re [\sigma _\mathrm {m}]=\Re [\sigma _{ii}]/2$ in the cylinder. The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$.

Download Full Size | PDF

3.2 Topology optimization for unidirectional and multidirectional cloaking

Topology optimization has been demonstrated for minimizing $F$ and optimal level set functions are explored by the CMA-ES under the perimeter constraint. Figure 3 shows the improvements of the fitness [Fig. 3(a)], the values of objective functions [Fig. 3(c)–(e)], and the perimeter [Fig. 3(f)] of the structure that exhibits the smallest $F$ from $\lambda$-candidate solutions in each generation $g$. CMA-ES generates candidate solutions based on the distribution parameters with random numbers in Eq. (14), then as shown in Fig. 3(a), the fitness value oscillates up and down slightly, but decreases as $g$ is updated. The optimization computation is considered to have converged when $\Delta \boldsymbol{m}_\mathrm {error}$ in Fig. 3(b) is less than or equal to 0.01, the calculation is terminated, and the structure that records the smallest $F$ in the process is the optimized structure. By imposing structural symmetries about not only $x$ and $y$ axes but also diagonal axes $y=\pm x$, multidirectional cloaking is also attempted to be design. In Figs. 3(c)–(e), all the value of objective functions reaches less than one (i.e. $\max \big ( \Psi _E, \Psi _H, \Psi _p \big )<1.00$) at generation $g=352$ ($\tau =0.01$), $g=336$ ($\tau =0.001$), $g=256$ ($\tau =0.01$ with $y=\pm x$ symmetries), and $g=241$ ($\tau =0.001$ with $y=\pm x$ symmetries) for the first time. The smaller regularization coefficient, $\tau =0.001$, prioritize the improvement of performances for cloaking over reducing the perimeter [Fig. 3(c)–(e)], and the performances have been improved even with less $g$ compared to the topology optimization with strict perimeter constraints, $\tau =0.01$, while the value of perimeter, implying the simplicity and good manufacturability of configuration, is not aggressively improved [Fig. 3(f)]. Because of the additional structural symmetry about $y=\pm x$, the number of design variables can be reduced by about half, which results in convergence with less $g$ [Fig. 3(b)].

 figure: Fig. 3.

Fig. 3. Histories of (a) the fitness and (b) the convergence error and objective functions of cloaking for (c) the $E_z$-polarized wave, (d) the $H_z$-polarized wave, and (e) the acoustic wave, and (f) the structural perimeter. The optimization process was judged to have converged at $\Delta \boldsymbol{m}_\mathrm {error} < 0.01$ (b).

Download Full Size | PDF

Figure 4 shows the results of topology optimizations minimizing fitness, Eq. (12). The numerical demonstration obtains an optimal design [Fig. 4(b)] that realizes both polarization-independent [Fig. 4(c)(d)] and acoustic [Fig. 4(e)] cloaking. With the weaker constraint $\tau =0.001$ prioritizing performance improvement [Fig. 4(f)–(j)], the three objective functions fall below $0.0467$ in value [Fig. 4(i)], corresponding to a reduction in scattering to $4.67\%$ of that for a bare obstacle present without a cloak [Fig. 2]. Nearly perfect performances for triple-wave cloaking capabilities are thus obtained numerically. By additionally imposing structural symmetries on $y=\pm x$ [Fig. 5], triple-wave cloaking capabilities are demonstrated for multidirectional wave propagations [Fig. 6(b,d,f)]. Several dips in the three objective functions [Fig. 6(b,d,f)] in the results obtained with $y=\pm x$ symmetry, that indicate the emergence of cloaking are observed not only when $\theta =-\pi, 0$, and $\pi$ but also when $\theta =-\pi /2$ and $\pi /2$. Because the level set functions are discretized on points of square cells as shown in Fig. 1(b), it is difficult to strictly impose higher $s$-fold discrete rotational symmetry ($s>8$) in the structure. It is not impossible to impose additional rotational symmetries through the linear interpolation of the level set function in the square cells, but a structure with a slightly broken symmetry appears. In further increasing the number of incident angles at which cloaking can be achieved, it is practical to impose symmetry about the diagonal axis $y=\pm x$ as shown in Fig. 5 on the structure and then compute the objective functions for several plane waves propagating at different incidence angles, and minimize all the objective functions simultaneously.

 figure: Fig. 4.

Fig. 4. (a,f) Transitions in structural topology exhibiting lowest $F$ at each $g$ during the optimization process, (b,g) optimal configuration of a biphysical cloak and its performance for (c,h) $E_z$-polarized wave, (d,i) $H_z$-polarized wave, and (e,j) an acoustic wave with (a)–(e) $\tau =0.01$ and (f)–(j) $\tau =0.001$. Structural perimeter in the optimal configuration becomes (b) $L = 37.4$ and (g) $L = 47.0$. The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$. The number of design variables is $n=3463$; the number of samplings is set to $\lambda =400$.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a,f) Transitions in structural topology exhibiting lowest $F$ at each $g$ with diagonal symmetries during the optimization process, (b,g) Optimal configuration of a multidirectional biphysical cloaks and their performances for (c,h) $E_z$-polarized wave, (d,i) $H_z$-polarized wave, and (e,j) acoustic wave with (a)–(e) $\tau =0.01$ and (f)–(j) $\tau =0.001$. Structural perimeter in the optimal configuration becomes (b) $L = 37.9$ and (g) $L = 72.2$. The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$. The number of design variables is $n=1748$; the number of samplings is set to $\lambda =400$.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Incident-angle dependences of cloaking performances of optimal configurations shown in [(a,c,e)] Fig. 4(b)(g) and [(b,d,f)] Fig. 5(b)(g) with $y=\pm x$ symmetries for (a,b) the $E_z$-polarized wave, (c,d) the $H_z$-polarized wave, and (e,f) the acoustic wave.

Download Full Size | PDF

Having discussed and demonstrated topology optimization for biphysical cloaks for electromagnetic and acoustic waves of the same wavelength in air, sound waves usually have wavelengths much longer than those of electromagnetic waves. Therefore, we examined whether a biphysical cloak can be designed that works with electromagnetic and sound waves of different wavelengths by the topology optimization. Assuming an incident sound wave of longer wavelength than that of the electromagnetic wave, results of topology optimization and frequency response for the cloaking performance of the optimal structure [Fig. 8] show that biphysical cloaking is achievable at least under the conditions we had investigated.

STL data for the optimal configurations in Fig. 4(b,g), Fig. 5(b,g) and Fig. 8(a)–(d) can be downloaded from Dataset 1, Ref. [42]. See Appendix B for details.

4. Discussion

We described topology-optimized biphysical cloaks that perform scattering cancelations for $E_z$- and $H_z$-polarized waves and acoustic wave. The frequency responses of the cloaking performances [Fig. 7] confirm the simultaneous cloaking of the three waves having the same wavelength in air despite the waves having different boundary conditions. We note that the normalized frequency is equal to the inverse of the normalized wavelength, e.g., $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=R_\mathrm {D}/\lambda _\mathrm {e}$, where $\lambda _\mathrm {e}$ is the wavelength of the electromagnetic wave in air.

 figure: Fig. 7.

Fig. 7. Frequency dependences of cloaking performances of optimal configurations shown in [(a,c,e)] Fig. 4(b)(g) and [(b,d,f)] Fig. 5(b)(g) with $y=\pm x$ symmetries for (a,b) the $E_z$-polarized wave, (c,d) the $H_z$-polarized wave, and (e,f) the acoustic wave.

Download Full Size | PDF

This optimization suppresses the scattering of the electromagnetic and acoustic waves independently; no physical interaction between the two waves, i.e., photon-phonon interaction, occurs [43,44]. How does the elastic vibration arising from the acoustic wave influence the cloaking performance for electromagnetic waves? We estimate the Doppler shift of the electromagnetic wave impinging on an optimized structure. The normalized shift in frequency is

$$\frac{\Delta\omega_\mathrm{e} R_\mathrm{D}}{2\pi c} =\left( f_\mathrm{e}-f_\mathrm{e}\frac{c-v}{c} \right)\frac{R_\mathrm{D}}{c} =\frac{\omega_\mathrm{e} R_\mathrm{D}}{2\pi c}\frac{v}{c},$$
where $f_\mathrm {e}$ denotes the frequency of the electromagnetic wave and $v$ the speed of the optimal structure for positive $x$. Assuming a simple harmonic oscillation of the optimal structure, $v=v_\mathrm {max} \cos (2\pi t/T)$, the displacement of an optimal structure is estimated to be
$$u=\displaystyle\int_0^{T/4} v_\mathrm{max} \cos(2\pi t/T)dt =v_\mathrm{max} T/(2\pi),$$
where $u$ denotes the maximum displacement of an optimum structure, $t$ the time satisfying $v=v_\mathrm {max}$ at $t=0$, $T$ the period of vibration corresponding to that of acoustic wave, and $v\leq v_\mathrm {max}=2\pi f_\mathrm {a} u$ with $f_\mathrm {a}$ the acoustic wave frequency. As an example, given the maximum displacement [see Fig. 5(e)], the normalized shift in frequency, Eq. (26), becomes
$$\frac{\omega_\mathrm{e} R_\mathrm{D}}{2\pi c}\frac{v}{c} \leq \frac{\omega_\mathrm{e} R_\mathrm{D}}{2\pi c} 2\pi \frac{f_\mathrm{a}}{f_\mathrm{e}} \frac{u_\mathrm{max}}{\lambda_\mathrm{e}} = 3.92\times10^{{-}6},$$
which is ignorable when compared with the frequency response for the cloaking performance at $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=2.5$ [Fig. 7(a)–d)]. Here, we assume the electromagnetic and acoustic waves in air have the same wavelength and use $f_\mathrm {a}/f_\mathrm {e}=\sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}} \big /c$.

5. Conclusions

We showed numerically that both electromagnetic and acoustic propagating waves can be manipulated to offer biphysical cloaking and a triple-wave cloaking capability exhibiting near-perfect cloaking performance of an isotropic and homogeneous material. Concerning the optimal configurations obtained, the scattering of both electromagnetic waves and acoustic waves were suppressed and biphysical cloaking for both wave types was achieved after topology optimization. Additionally, the presented method addressed to incorporate some physical attributes in the structural optimization such as polarization for electromagnetic waves and material elasticity for sound waves. The effect of the interaction between electromagnetic and acoustic waves on the cloaking performance is limited and negligible. The performances of our optimal biphysical cloaks as well as the electromagnetic cloaks [45,46] are expected to be demonstrated in future experiments. Our method may be extended to design other biphysical devices by replacing an objective function with an appropriate function that evaluates device performance.

Appendix A

Acoustic performance of the topology-optimized electromagnetic cloak

Figures 7(e,f) and 8(i,j) show that the objective function for acoustic waves at frequencies not taken into account in the optimization is higher than that for electromagnetic waves. It is interesting to see what happens if topology optimizations are demonstrated for infimizing only the objective functions in electromagnetic cloaking and to see how the resulting device will perform for acoustic waves. Additionally, we can compare the results obtained with and without considering acoustic waves in our topology optimizations and make sure that our optimization approach does not ignore that term.

 figure: Fig. 8.

Fig. 8. Optimal configurations for different acoustic wave frequencies: $\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}})$ equal to (a) $0.5$, (b) $1.0$, (c) $1.5$, and (d) $2.0$, for fixed frequency of the electromagnetic wave, $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=2.5$ and $\tau =0.01$. (a)–(d) The number of design variables is $n=1748$; the number of samplings is set to $\lambda =400$. Frequency responses of the cloaking performances for (e,h) $E_z$-polarized electromagnetic wave, (f,i) $H_z$-polarized electromagnetic wave and (g,j) acoustic cloaks.

Download Full Size | PDF

We then investigate how the topology-optimized electromagnetic cloak, designed without considering the objective function for acoustic cloaking, performs for acoustic waves. A topology optimization has been demonstrated for a polarization-independent cloak by infimizing the fitness without the objective function for acoustic cloaking:

$$\underset{\phi}{\mathrm{inf}}\qquad F_{EH}= \max\big( \Psi_E, \Psi_H, \tau L \big).$$

We also investigate the cloaking performance of the acoustic cloak for electromagnetic waves presented in Ref. [29], for which the fitness is infimized as

$$\underset{\phi}{\mathrm{inf}}\qquad F_{p}= \max\big( \Psi_p, \tau L \big).$$

Figures 9(a)–(d) show the optimization results of infimizing $F_{EH}$ without incorporating $\Psi _p$ whereas Figs. 9(e)–(h) show those of infimizing $F_{p}$ without incorporating $\Psi _E$ and $\Psi _H$ obtained from Ref. [29]. These results show that the unconsidered objective function performs poorly as shown in Fig. 9(d) and Fig. 9(f,g), and we confirm that the our optimization approach does not ignore every single objective function incorporated in the fitness $F$ in Eq. (12).

 figure: Fig. 9.

Fig. 9. (a) Polarization-independent electromagnetic cloak obtained by infimizing $F_{EH}$ and its cloaking performance for (b) the $E_z$-polarized wave, (c) the $H_z$-polarized wave, and (d) the acoustic wave. (e) Topology-optimized acoustic cloak obtained by infimizing $F_{p}$ in Ref. [29] and its cloaking performance for (f) the $E_z$-polarized wave, (g) the $H_z$-polarized wave, and (h) the acoustic wave. The regularization coefficient is set at $\tau =0.001$ and diagonal structural symmetries are imposed. (a)–(d) The number of design variables is $n=1748$; the number of samplings is set to $\lambda =400$. (e)–(h) The number of design variables is $n=2009$; the number of samplings is set to $\lambda =160$ in Ref. [29]. (e,h) Reprinted from [29] with the permission of AIP Publishing.

Download Full Size | PDF

Appendix B

STL data of the presented optimal configurations

We provide stereolithography (STL) data of the optimal configurations shown in Fig. 4(b,g), Fig. 5(b,g), and Fig. 8(a)–(d) for experimental demonstrations of the biphysical cloaks designed through topology optimization as we show in Dataset 1, Ref. [42]. Optimal structures made from ABS resin are provided as shown in Fig. 10. Each STL file name corresponds to the figure number of optimal configurations in the manuscript; e.g., STL of the optimal configuration in Fig. 4(b) is denoted “Fig 4b.stl".

 figure: Fig. 10.

Fig. 10. (a) Optimal configuration of electromagnetic-acoustic biphysical cloak in Fig. 5(b) in the manuscript. (b) Corresponding STL file named “Fig5b.stl”.

Download Full Size | PDF

In the optimizations in Figs. 4 and 5, we assume that the wavelengths of the electromagnetic and acoustic waves are $\lambda _\mathrm {e}=125$mm and $\lambda _\mathrm {a}=125$mm respectively and that the frequencies of electromagnetic and acoustic waves to be approximately $f_\mathrm {e}=2.40$GHz and $f_\mathrm {a}=2.72$kHz respectively. The radius of the fixed design domain becomes $R_\mathrm {D}=312.5$mm. The height in the $z$ direction is set at $20$ mm. We keep in mind that when the aluminum cylinder is in contact with the ABS, the two structures are simulated as adhering to each other. Operating wavelengths and frequencies are listed in Table 2.

Tables Icon

Table 2. Optimum Structures, STL Data, Operating Wavelengths, and Frequency.

Roughly speaking, the smallest feature becomes the discretized cell size $L_\mathrm {grid}$ estimated from the wavelength, $\lambda _\mathrm {e}$, and radius of the fixed design domain, $R_\mathrm {D}$. The ratio between $\lambda _\mathrm {e}$ and $R_\mathrm {D}$ is obtained from the normalized frequency as $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)= R_\mathrm {D}/\lambda _\mathrm {e}=2.5$. The cell size is then

$$L_\mathrm{grid}=R_\mathrm{D}/70=\lambda_\mathrm{e}\times2.5/70=125 \mathrm{mm}\times2.5/70\approx 4.46\mathrm{mm}.$$

The smallest size of a structural feature is estimated as being approximately 4.46 mm. Meanwhile, by regarding $L_\mathrm {grid}$ as the fabrication tolerance of the patterned ABS, the wavelength is inversely calculated as

$$\lambda_\mathrm{e}=L_\mathrm{grid}\times 70/2.5=L_\mathrm{grid}\times 28.$$

If we assume that the manufacturing tolerance of recent 3D printers is $10\mu$m–$25\mu$m, the wavelength $\lambda _\mathrm {e}$ becomes $280\mu$m–$700\mu$m. We note that this minimal structural feature can be improved by increasing the regularization coefficient $\tau$, as described in this paper. If the minimum structural features are desired to be sufficiently larger than $L_\mathrm {grid}$, it is recommended to set the coefficient $\tau$ larger.

Acknowledgments

We thank Richard Haase, Ph.D., from Edanz (https://jp.edanz.com/ac) for editing a draft of this manuscript.

Disclosures

The authors declare that there are no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request. STL data for the optimal configurations in Fig. 4(b,g), Fig. 5(b,g), and Fig. 8(a)–(d) can be downloaded from Dataset 1, Ref. [42] for supporting content.

References

1. J. Y. Li, Y. Gao, and J. P. Huang, “A bifunctional cloak using transformation media,” J. Appl. Phys. 108(7), 074504 (2010). [CrossRef]  

2. M. Moccia, G. Castaldi, S. Savo, Y. Sato, and V. Galdi, “Independent manipulation of heat and electrical current via bifunctional metamaterials,” Phys. Rev. X 4(2), 021025 (2014). [CrossRef]  

3. Y. Ma, Y. Liu, M. Raza, Y. Wang, and S. He, “Experimental demonstration of a multiphysics cloak: manipulating heat flux and electric current simultaneously,” Phys. Rev. Lett. 113(20), 205501 (2014). [CrossRef]  

4. C. Lan, K. Bi, Z. Gao, B. Li, and J. Zhou, “Achieving bifunctional cloak via combination of passive and active schemes,” Appl. Phys. Lett. 109(20), 201903 (2016). [CrossRef]  

5. T. Yang, X. Bai, D. Gao, L. Wu, B. Li, J. T. L. Thong, and C.-W. Qiu, “Invisible sensors: simultaneous sensing and camouflaging in multiphysical fields,” Adv. Mater. 27(47), 7752–7758 (2015). [CrossRef]  

6. C. Lan, K. Bi, X. Fu, B. Li, and J. Zhou, “Bifunctional metamaterials with simultaneous and independent manipulation of thermal and electric fields,” Opt. Express 24(20), 23072–23080 (2016). [CrossRef]  

7. G. Fujii and Y. Akimoto, “Optimizing the structural topology of bifunctional invisible cloak manipulating heat flux and direct current,” Appl. Phys. Lett. 115(17), 174101 (2019). [CrossRef]  

8. V. G. Veselago, “The electrodynamics of substances with simultaneously negative values of ε and μ,” Sov. Phys. Usp. 10(4), 509–514 (1968). [CrossRef]  

9. J. B. Pendry, A. J. Holden, D. J. Robbins, and W. J. Stewart, “Magnetism from conductors and enhanced nonlinear phenomena,” IEEE Trans. Microw. Theory Tech. 47(11), 2075–2084 (1999). [CrossRef]  

10. U. Leonhardt, “Optical conformal mapping,” Science 312(5781), 1777–1780 (2006). [CrossRef]  

11. J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic fields,” Science 312(5781), 1780–1782 (2006). [CrossRef]  

12. S. A. Cummer and D. Schurig, “One path to acoustic cloaking,” New J. Phys. 9(3), 45 (2007). [CrossRef]  

13. J. Xu, X. Jiang, N. Fang, E. Georget, R. Abdeddaim, J.-M. Geffrin, M. Farhat, P. Sabouroux, S. Enoch, and S. Guenneau, “Molding acoustic, electromagnetic and water waves with a single cloak,” Sci. Rep. 5(1), 10678 (2015). [CrossRef]  

14. Y. Yang, H. Wang, F. Yu, Z. Xu, and H. Chen, “A metasurface carpet cloak for electromagnetic, acoustic and water waves,” Sci. Rep. 6(1), 20219 (2016). [CrossRef]  

15. F. Sun, Y. Liu, and S. He, “Surface transformation multi-physics for controlling electromagnetic and acoustic waves simultaneously,” Opt. Express 28(1), 94–106 (2020). [CrossRef]  

16. M. Zandehshahvar, Y. Kiarashi, M. Zhu, H. Maleki, T. Brown, and A. Adibi, “Manifold learning for knowledge discovery and intelligent inverse design of photonic nanostructures: Breaking the geometric complexity,” arXiv p. 2102.04454 (2021).

17. M. Zandehshahvar, Y. Kiarashi, M. Chen, R. Barton, and A. Adibi, “Inverse design of photonic nanostructures using dimensionality reduction: reducing the computational complexity,” Opt. Lett. 46(11), 2634–2637 (2021). [CrossRef]  

18. P. R. Wiecha, A. Arbouet, C. Girard, and O. L. Muskens, “Deep learning in nano-photonics: inverse design and beyond,” Photonics Res. 9(5), B182–B200 (2021). [CrossRef]  

19. G. Fujii, Y. Akimoto, and M. Takahashi, “Exploring optimal topology of thermal cloaks by CMA-ES,” Appl. Phys. Lett. 112(6), 061108 (2018). [CrossRef]  

20. G. Fujii, Y. Akimoto, and M. Takahashi, “Direct-current electric invisibility through topology optimization,” J. Appl. Phys. 123(23), 233102 (2018). [CrossRef]  

21. J. Andkjaer and O. Sigmund, “Topology optimized low-contrast all-dielectric optical cloak,” Appl. Phys. Lett. 98(2), 021112 (2011). [CrossRef]  

22. J. Andkjaer, N. A. Mortensen, and O. Sigmund, “Towards all-dielectric, polarization-independent optical cloaks,” Appl. Phys. Lett. 100(10), 101106 (2012). [CrossRef]  

23. G. Fujii, H. Watanabe, T. Yamada, T. Ueta, and M. Mizuno, “Level set based topology optimization for optical cloaks,” Appl. Phys. Lett. 102(25), 251106 (2013). [CrossRef]  

24. G. Fujii and T. Ueta, “Topology-optimized carpet cloaks based on a level-set boundary expression,” Phys. Rev. E 94(4), 043301 (2016). [CrossRef]  

25. N. Kishimoto, K. Izui, S. Nishiwaki, and T. Yamada, “Optimal design of electromagnetic cloaks with multiple dielectric materials by topology optimization,” Appl. Phys. Lett. 110(20), 201104 (2017). [CrossRef]  

26. C. Gustavo Méndez, J. M. Podestá, O. Lloberas-Valls, S. Toro, A. E. Huespe, and J. Oliver, “Computational material design for acoustic cloaking,” Int. J. Numer. Methods Engrg. 112(10), 1353–1380 (2017). [CrossRef]  

27. Z. Ma, O. Stalnov, and X. Huang, “Design method for an acoustic cloak in flows by topology optimization,” Acta Mech. Sin. 35(5), 964–971 (2019). [CrossRef]  

28. P. Chen, M. R. Haberman, and O. Ghattas, “Optimal design of acoustic metamaterial cloaks under uncertainty,” J. Comput. Phys. 431, 110114 (2021). [CrossRef]  

29. G. Fujii, M. Takahashi, and Y. Akimoto, “Acoustic cloak designed by topology optimization for acoustic-elastic coupled systems,” Appl. Phys. Lett. 118(10), 101102 (2021). [CrossRef]  

30. N. Hansen and A. Ostermeier, “Completely derandomized self-adaptation in evolution strategies,” Evol. Comput. 9(2), 159–195 (2001). [CrossRef]  

31. G. Fujii, M. Takahashi, and Y. Akimoto, “CMA-ES-based structural topology optimization using a level set boundary expression—Application to optical and carpet cloaks,” Comput. Methods Appl. Mech. Eng. 332, 624–643 (2018). [CrossRef]  

32. N. Hansen and A. Auger, Principled design of continuous stochastic search: from theory to practice (Springer, 2014).

33. L. M. Rios and N. V. Sahinidis, “Derivative-free optimization: a review of algorithms and comparison of software implementations,” J. Glob. Optim. 56(3), 1247–1293 (2013). [CrossRef]  

34. G. Fujii and Y. Akimoto, “DC carpet cloak designed by topology optimization based on covariance matrix adaptation evolution strategy,” Opt. Lett. 44(8), 2057–2060 (2019). [CrossRef]  

35. G. Fujii and Y. Akimoto, “dc electric cloak concentrator via topology optimization,” Phys. Rev. E 102(3), 033308 (2020). [CrossRef]  

36. G. Fujii and Y. Akimoto, “Cloaking a concentrator in thermal conduction via topology optimization,” Int. J. Heat Mass Transfer 159, 120082 (2020). [CrossRef]  

37. N. Hansen, A. S. P. Niederberger, L. Guzzella, and P. Koumoutsakos, “A method for handling uncertainty in evolutionary optimization with an application to feedback control of combustion,” IEEE Trans. Evol. Comput. 13(1), 180–197 (2009). [CrossRef]  

38. N. Hansen, “The CMA evolution strategy: a tutorial,” arXiv: p. 1604.00772 (2016).

39. N. Hansen and S. Kern, “Evaluating the CMA evolution strategy on multimodal test functions,” Parallel Problem Solving from Nature - PPSN VIII pp. 282–291 (2004).

40. Y. Akimoto and N. Hansen, “Diagonal acceleration for covariance matrix adaptation evolution strategies,” Evolutionary Computation 28(3), 405–435 (2020). [CrossRef]  

41. G. Fujii and Y. Akimoto, “Topology-optimized thermal carpet cloak expressed by an immersed-boundary level-set method via a covariance matrix adaptation evolution strategy,” Int. J. Heat Mass Transfer 137, 1312–1322 (2019). [CrossRef]  

42. G. Fujii, “STL data of topology-optimized biphysical cloaks manipulating electromagnetic and acoustic waves,” figshare (2022), https://doi.org/10.6084/m9.figshare.19030130.

43. Y. Tanaka, J. Upham, T. Nagashima, T. Sugiya, T. Asano, and S. Noda, “Dynamic control of the q factor in a photonic crystal nanocavity,” Nat. Mater. 6(11), 862–865 (2007). [CrossRef]  

44. T. Ueta, “Photon-phonon interaction in photonic crystals,” IOP Conf. Ser.: Mater. Sci. Eng. 10, 012103 (2010). [CrossRef]  

45. G. Fujii, H. Watanabe, T. Yamada, T. Ueta, and M. Mizuno, “Level set based topology optimization for optical cloaks containing a large scattering object,” in Proceedings of WCSMO10, (2013), p. 5283.

46. L. Lan, F. Sun, Y. Liu, C. K. Ong, and Y. Ma, “Experimentally demonstrated a unidirectional electromagnetic cloak designed by topology optimization,” Appl. Phys. Lett. 103(12), 121113 (2013). [CrossRef]  

Supplementary Material (1)

NameDescription
Dataset 1       STL data of topology-optimized biphysical cloaks manipulating electromagnetic and acoustic waves

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request. STL data for the optimal configurations in Fig. 4(b,g), Fig. 5(b,g), and Fig. 8(a)–(d) can be downloaded from Dataset 1, Ref. [42] for supporting content.

42. G. Fujii, “STL data of topology-optimized biphysical cloaks manipulating electromagnetic and acoustic waves,” figshare (2022), https://doi.org/10.6084/m9.figshare.19030130.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. Schematic of a topology optimization for an electromagnetic-acoustic biphysical cloak. Domain sizes are $R_\mathrm {obs}=R_\mathrm {D}/3$ , $L_x=R_\mathrm {D}\times 2$ , $L_y=R_\mathrm {D}\times 3/2$ , and $L_\mathrm {grid}=R_\mathrm {D}/70$ .
Fig. 2.
Fig. 2. (a) Bare cylinder without a cloak and its scattering of (b) $E_z$ -polarized wave, (c) $H_z$ -polarized wave, and (d) acoustic wave around the cylinder; (d) mean stress $\Re [\sigma _\mathrm {m}]=\Re [\sigma _{ii}]/2$ in the cylinder. The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$ .
Fig. 3.
Fig. 3. Histories of (a) the fitness and (b) the convergence error and objective functions of cloaking for (c) the $E_z$ -polarized wave, (d) the $H_z$ -polarized wave, and (e) the acoustic wave, and (f) the structural perimeter. The optimization process was judged to have converged at $\Delta \boldsymbol{m}_\mathrm {error} < 0.01$ (b).
Fig. 4.
Fig. 4. (a,f) Transitions in structural topology exhibiting lowest $F$ at each $g$ during the optimization process, (b,g) optimal configuration of a biphysical cloak and its performance for (c,h) $E_z$ -polarized wave, (d,i) $H_z$ -polarized wave, and (e,j) an acoustic wave with (a)–(e) $\tau =0.01$ and (f)–(j) $\tau =0.001$ . Structural perimeter in the optimal configuration becomes (b) $L = 37.4$ and (g) $L = 47.0$ . The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$ . The number of design variables is $n=3463$ ; the number of samplings is set to $\lambda =400$ .
Fig. 5.
Fig. 5. (a,f) Transitions in structural topology exhibiting lowest $F$ at each $g$ with diagonal symmetries during the optimization process, (b,g) Optimal configuration of a multidirectional biphysical cloaks and their performances for (c,h) $E_z$ -polarized wave, (d,i) $H_z$ -polarized wave, and (e,j) acoustic wave with (a)–(e) $\tau =0.01$ and (f)–(j) $\tau =0.001$ . Structural perimeter in the optimal configuration becomes (b) $L = 37.9$ and (g) $L = 72.2$ . The normalized frequency is set to $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}}) =2.5$ . The number of design variables is $n=1748$ ; the number of samplings is set to $\lambda =400$ .
Fig. 6.
Fig. 6. Incident-angle dependences of cloaking performances of optimal configurations shown in [(a,c,e)] Fig. 4(b)(g) and [(b,d,f)] Fig. 5(b)(g) with $y=\pm x$ symmetries for (a,b) the $E_z$ -polarized wave, (c,d) the $H_z$ -polarized wave, and (e,f) the acoustic wave.
Fig. 7.
Fig. 7. Frequency dependences of cloaking performances of optimal configurations shown in [(a,c,e)] Fig. 4(b)(g) and [(b,d,f)] Fig. 5(b)(g) with $y=\pm x$ symmetries for (a,b) the $E_z$ -polarized wave, (c,d) the $H_z$ -polarized wave, and (e,f) the acoustic wave.
Fig. 8.
Fig. 8. Optimal configurations for different acoustic wave frequencies: $\omega _\mathrm {a} R_\mathrm {D}/(2\pi \sqrt {\kappa _\mathrm {a}/\rho _\mathrm {a}})$ equal to (a) $0.5$ , (b) $1.0$ , (c) $1.5$ , and (d) $2.0$ , for fixed frequency of the electromagnetic wave, $\omega _\mathrm {e} R_\mathrm {D}/(2\pi c)=2.5$ and $\tau =0.01$ . (a)–(d) The number of design variables is $n=1748$ ; the number of samplings is set to $\lambda =400$ . Frequency responses of the cloaking performances for (e,h) $E_z$ -polarized electromagnetic wave, (f,i) $H_z$ -polarized electromagnetic wave and (g,j) acoustic cloaks.
Fig. 9.
Fig. 9. (a) Polarization-independent electromagnetic cloak obtained by infimizing $F_{EH}$ and its cloaking performance for (b) the $E_z$ -polarized wave, (c) the $H_z$ -polarized wave, and (d) the acoustic wave. (e) Topology-optimized acoustic cloak obtained by infimizing $F_{p}$ in Ref. [29] and its cloaking performance for (f) the $E_z$ -polarized wave, (g) the $H_z$ -polarized wave, and (h) the acoustic wave. The regularization coefficient is set at $\tau =0.001$ and diagonal structural symmetries are imposed. (a)–(d) The number of design variables is $n=1748$ ; the number of samplings is set to $\lambda =400$ . (e)–(h) The number of design variables is $n=2009$ ; the number of samplings is set to $\lambda =160$ in Ref. [29]. (e,h) Reprinted from [29] with the permission of AIP Publishing.
Fig. 10.
Fig. 10. (a) Optimal configuration of electromagnetic-acoustic biphysical cloak in Fig. 5(b) in the manuscript. (b) Corresponding STL file named “Fig5b.stl”.

Tables (2)

Tables Icon

Table 1. Material Properties in the Optimization.

Tables Icon

Table 2. Optimum Structures, STL Data, Operating Wavelengths, and Frequency.

Equations (32)

Equations on this page are rendered with MathJax. Learn more.

Ψ X = 1 Ψ X n Ω o u t | X s | 2 d Ω ,
Ψ X n = Ω o u t | X b a r e s | 2 d Ω ,
E , i i + ω e 2 ϵ c 2 E = 0 i n   Ω D Ω s , Ω s , Ω o u t ,
ϵ 1 H , i i + ω e 2 c 2 H = 0 i n   Ω D Ω s , Ω s , Ω o u t ,
E = 0 o n   Γ o b s s , Γ D o b s ,
n i P E C H , i = 0 o n   Γ o b s s , Γ D o b s ,
ρ a 1 p , i i + ω a 2 κ a 1 p = 0 i n   Ω D Ω s , Ω o u t ,
σ j i , j + ρ e ω a 2 u i = 0 i n   Ω s , Ω o b s ,
p , i n i s = ω a 2 ρ a u i n i s o n   Γ D s , Γ D o b s ,
σ j i ( u ) n i s = p n i s o n   Γ D s , Γ D o b s ,
Ω D Ω s = { x   |   1 ϕ ( x ) < 0 } , Γ D s = { x   |   ϕ ( x ) = 0 } , Ω s Γ D s = { x   |   0 < ϕ ( x ) 1 } .
i n f ϕ F = max ( Ψ E , Ψ H , Ψ p , τ L ) ,
m ( 0 ) = 0.5 ( b u + b l ) 1 n , σ ( 0 ) = 0.3 ( b u b l ) , C ( 0 ) = I n , p σ ( 0 ) = 0 , p C ( 0 ) = 0 ,
ϕ i ( g + 1 ) = m ( g ) + σ ( g ) B D z i ( g + 1 ) ,
C ( g ) = B D 2 B .
y i : λ ( g + 1 ) = ( ϕ i : λ ( g + 1 ) m ( g ) ) / σ ( g ) .
m ( g + 1 ) = m ( g ) + σ ( g ) m y ,
σ ( g + 1 ) = σ ( g )   exp ( c σ d σ ( | | p σ ( g + 1 ) | | E [ | | N ( 0 , I n ) | | ] 1 ) ) ,
C ( g + 1 ) = C ( g ) + h σ ( g + 1 ) c 1 ( p C ( g + 1 ) ( p C ( g + 1 ) ) C ( g ) ) + c μ i = 1 λ w i ( y i : λ ( g + 1 ) ( y i : λ ( g + 1 ) ) C ( g ) ) ,
p C ( g + 1 ) = ( 1 c C ) p C ( g ) + h σ ( g + 1 ) c C ( 2 c C ) μ e f f m y ,
p σ ( g + 1 ) = ( 1 c σ ) p σ ( g ) + c σ ( 2 c σ ) μ e f f C ( g ) 1 m y ,
C ( g ) 1 = B D 1 B .
m y = i = 1 μ w i y i : λ ( g + 1 ) ,
h σ ( g + 1 ) = { 1   | | p σ ( g + 1 ) | | < ( 1.5 + 1 n 0.5 ) E [ | | N ( 0 , I n ) | | ] , 0   o t h e r w i s e ,
Δ m e r r o r = | | m ( g ) m ( g 1 ) | | | | m ( g ) | | 0.01.
Δ ω e R D 2 π c = ( f e f e c v c ) R D c = ω e R D 2 π c v c ,
u = 0 T / 4 v m a x cos ( 2 π t / T ) d t = v m a x T / ( 2 π ) ,
ω e R D 2 π c v c ω e R D 2 π c 2 π f a f e u m a x λ e = 3.92 × 10 6 ,
i n f ϕ F E H = max ( Ψ E , Ψ H , τ L ) .
i n f ϕ F p = max ( Ψ p , τ L ) .
L g r i d = R D / 70 = λ e × 2.5 / 70 = 125 m m × 2.5 / 70 4.46 m m .
λ e = L g r i d × 70 / 2.5 = L g r i d × 28.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.