
This paper presents a numerical method for solving the inverse problem of reconstructing the shape of periodic structures from scattering data. This inverse problem is motivated by applications in the nondestructive evaluation of photonic crystals. The numerical method belongs to the class of sampling methods that aim to construct an imaging function for the shape of the periodic structure using scattering data. By extending the results of Nguyen, Stahl, and Truong [Inverse Problems, 39:065013, 2023], we studied a simple imaging function that uses half the data required by the numerical method in the cited paper. Additionally, this imaging function is fast, simple to implement, and very robust against noise in the data. Both isotropic and anisotropic cases were investigated, and numerical examples were presented to demonstrate the performance of the numerical method.
Citation: John Daugherty, Nate Kaduk, Elena Morgan, Dinh-Liem Nguyen, Peyton Snidanko, Trung Truong. On fast reconstruction of periodic structures with partial scattering data[J]. Electronic Research Archive, 2024, 32(11): 6481-6502. doi: 10.3934/era.2024303
[1] | Xinlin Cao, Huaian Diao, Jinhong Li . Some recent progress on inverse scattering problems within general polyhedral geometry. Electronic Research Archive, 2021, 29(1): 1753-1782. doi: 10.3934/era.2020090 |
[2] | Yujie Wang, Enxi Zheng, Wenyan Wang . A hybrid method for the interior inverse scattering problem. Electronic Research Archive, 2023, 31(6): 3322-3342. doi: 10.3934/era.2023168 |
[3] | Deyue Zhang, Yukun Guo . Some recent developments in the unique determinations in phaseless inverse acoustic scattering theory. Electronic Research Archive, 2021, 29(2): 2149-2165. doi: 10.3934/era.2020110 |
[4] | Yao Sun, Lijuan He, Bo Chen . Application of neural networks to inverse elastic scattering problems with near-field measurements. Electronic Research Archive, 2023, 31(11): 7000-7020. doi: 10.3934/era.2023355 |
[5] | Yan Chang, Yukun Guo . Simultaneous recovery of an obstacle and its excitation sources from near-field scattering data. Electronic Research Archive, 2022, 30(4): 1296-1321. doi: 10.3934/era.2022068 |
[6] | Wenya Shi, Xinpeng Yan, Zhan Huan . Faster free pseudoinverse greedy block Kaczmarz method for image recovery. Electronic Research Archive, 2024, 32(6): 3973-3988. doi: 10.3934/era.2024178 |
[7] | Youjun Deng, Hongyu Liu, Xianchao Wang, Dong Wei, Liyan Zhu . Simultaneous recovery of surface heat flux and thickness of a solid structure by ultrasonic measurements. Electronic Research Archive, 2021, 29(5): 3081-3096. doi: 10.3934/era.2021027 |
[8] | Hyungyeong Jung, Sunghwan Moon . Reconstruction of the initial function from the solution of the fractional wave equation measured in two geometric settings. Electronic Research Archive, 2022, 30(12): 4436-4446. doi: 10.3934/era.2022225 |
[9] | Qing Tian, Heng Zhang, Shiyu Xia, Heng Xu, Chuang Ma . Cross-view learning with scatters and manifold exploitation in geodesic space. Electronic Research Archive, 2023, 31(9): 5425-5441. doi: 10.3934/era.2023275 |
[10] | Jiyu Sun, Jitao Zhang . Muti-frequency extended sampling method for the inverse acoustic source problem. Electronic Research Archive, 2023, 31(7): 4216-4231. doi: 10.3934/era.2023214 |
This paper presents a numerical method for solving the inverse problem of reconstructing the shape of periodic structures from scattering data. This inverse problem is motivated by applications in the nondestructive evaluation of photonic crystals. The numerical method belongs to the class of sampling methods that aim to construct an imaging function for the shape of the periodic structure using scattering data. By extending the results of Nguyen, Stahl, and Truong [Inverse Problems, 39:065013, 2023], we studied a simple imaging function that uses half the data required by the numerical method in the cited paper. Additionally, this imaging function is fast, simple to implement, and very robust against noise in the data. Both isotropic and anisotropic cases were investigated, and numerical examples were presented to demonstrate the performance of the numerical method.
We consider the inverse scattering problem for two-dimensional periodic structures which are assumed to be unboundedly periodic in the horizontal direction and bounded in the vertical direction. These periodic strutures are motivated by one-dimensional photonic crystals [1]. The inverse scattering problem of interest aims to reconstruct the shape of these periodic structures from boundary scattering data. This inverse problem is motivated by applications of nondestructive testing for photonic crystals.
During the past two decades, there has been a considerable amount of research on numerical methods for this inverse problem, see [2,3,4,5,6,7,8,9,10,11]. A significant portion of these studies involves the factorization method [12]. This method, which belongs to the class of sampling or qualitative methods, was introduced by D. Colton and A. Kirsch [13,14]. The factorization method aims to construct a necessary and sufficient characterization of the unknown scatterer from multi-static scattering data. It is a fast and non-iterative method that does not require advanced a priori information about the unknown scatterers. However, it is not very robust against noise in the data when imaging periodic structures [2]. For periodic scattering structures modeled by a smooth periodic function multiplied by a small surface deformation parameter, the near-field imaging method [3,15] can provide reconstruction with super resolution.
Inspired by the direct or orthogonality sampling methods [16,17,18,19], a sampling method with novel imaging functions has been recently developed in [20,21] for imaging periodic structures from scattering data. Similar to the direct sampling method, this new sampling method is fast, stable, and simple to implement, avoiding the need to solve an ill-posed problem. Numerical studies also show that this method is more accurate than the direct sampling method and more stable than the factorization method for imaging periodic structures. This paper extends the results in [20]. Specifically, we modify the imaging function developed in [20] to use only half the data required by the original imaging function. Additionally, this imaging function is fast, simple to implement, and very robust against noise in the data. We study the modified imaging function for both isotropic and anisotropic cases of periodic scattering media. Numerical examples are presented to demonstrate the performance of the numerical method and to compare it with the direct sampling method and the factorization method.
The paper is organized as follows. The basics of the scattering from periodic media and the inverse problem of interest are described in Section 2. The modified imaging function and its analysis for the isotropic case are discussed in Section 3. Results for the case of anisotropic media are presented in Section 4. Finally, Section 5 is dedicated to a numerical study of the sampling method and its comparison to the factorization method and the orthogonality sampling method.
We consider a two-dimensional medium that is 2π periodic in the x1-direction and bounded in the x2-direction. While we choose the period to be 2π for convenience, the medium can have any arbitrary period. For α∈R, a function f:R2→C is α-quasiperiodic in x1 if
f(x1+2πj,x2)=ei2πjαf(x1,x2),∀j∈Z. |
Let q:R2→R be a bounded function that represents the material parameter of the medium relative to that of the background. Assume that q is 2π-periodic in x1, supp(q) is bounded in x2, and q=0 outside of the medium. We consider an α-quasiperiodic incident field, uin, to illuminate the medium. The medium scatters the incident field and produces a scattered field, usc, which satisfies the Helmholtz equation,
Δusc+k2usc=−k2quin R2, | (2.1) |
where u:=usc+uin is the total field, and k>0 is the wave number. It is well known that usc is also α-quasiperiodic in x1. Thus, the problem can be reduced to one period:
Ω:=(−π,π)×R. |
Let D:=supp(q)∩Ω and h>0 such that
h>sup{|x2|:x=(x1,x2)∈D}. | (2.2) |
To ensure that the direct scattering problem is well-posed, we impose the Rayleigh radiation condition for the scattered field as
usc(x)={∑j∈Zu+jeiαjx1+iβj(x2−h),x2≥h,∑j∈Zu−jeiαjx1−iβj(x2+h),x2≤−h, | (2.3) |
where
αj:=α+j,βj:={√k2−α2j,k2>α2ji√α2j−k2,k2<α2jj∈Z, |
and u±j, j∈Z, are called the Rayleigh coefficients of the scattered field. The condition (2.3) means that usc is an outgoing wave. Note that we exclude the case where k=αj for some j, which is known as Wood's anomaly. See [22] for a detailed discussion on well-posedness of the direct problem under some assumption on q and wave number k. For the study of the inverse problem of this paper, we will assume that the direct problem (2.1)–(2.3) is well-posed. Let
Γ±h:=(−π,π)×{±h}, |
and the inverse problem of interest can be stated as follows.
Inverse problem: Using multiple incident fields at a fixed wave number k to illuminate an unknown scattering medium, determine the geometry D of the medium from the measurement of the scattered fields on either Γ+h or Γ−h.
To solve this inverse problem, we introduce a new imaging function, which is inspired by that for the case of full measurement on Γ+h∪Γ−h in [20].
For N∈N and l=1,2,…,N, we denote by uin(x,l) the incident fields used to illuminate the unknown medium. The corresponding scattered fields and total fields are denoted by usc(x,l) and u(x,l), respectively. It is well-known that the scattered fields satisfy the Lippmann-Schwinger integral equation [23]:
usc(x,l)=k2∫DG(x,y)q(y)u(y,l)dy, | (3.1) |
where G(x,y) is the α-quasiperiodic Green's function, which admits the following series representation [1]:
G(x,y):=i4π∑j∈Z1βjeiαj(x1−y1)+iβj|x2−y2|,x,y∈Ω,x≠y. | (3.2) |
Note that G(⋅,y) also admits a Rayleigh expansion similar to (2.3), and its Rayleigh coefficients g±j(y) are given by
g±j(y)=i4πβje−iαjy1∓iβj(y2∓h). | (3.3) |
Now, let us discuss the following lemma, which is the motivation behind the definition of the imaging function.
Lemma 1. Let Ωh:=(−π,π)×(−h,h) and
F(x,y):=G(x,y)−¯G(y,x)2i. |
Then, for x∈Γ+h∪Γ−h and y∈Ωh,
F(x,y)=18π2∑j:βj>01βjeiαj(x1−y1)cos(βj(x2−y2)). |
Moreover, for y,z∈Ωh,
∫Γ+h∪Γ−h∂F(x,y)∂ν(x)¯G(x,z)−F(x,y)∂¯G(x,z)∂ν(x) ds(x)=F(z,y), | (3.4) |
where ν(x) denotes the normal vector.
Proof. For x∈Γ+h and y∈Ωh, we have x2>y2 and
G(x,y)=i4π∑j∈Z1βjeiαj(x1−y1)+iβj(x2−y2), | (3.5) |
¯G(y,x)=−i4π∑j∈Z1¯βje−iαj(y1−x1)−i¯βj(x2−y2). |
From these two expressions, we can see that the terms for which βj is complex-valued in G(x,y) coincide with their counterparts in ¯G(y,x). Thus,
F(x,y)=18π∑j:βj>01βjeiαj(x1−y1)(eiβj(x2−y2)+e−iβj(x2−y2))=14π∑j:βj>01βjeiαj(x1−y1)cos(βj(x2−y2)). |
Next, for fixed z,y∈Ωh,
∫Γ+h∂F(x,y)∂ν¯G(x,z)−F(x,y)∂¯G(x,z)∂ν ds(x)=i16π2∫Γ+h∑j:βj>0eiαj(x1−y1)sin(βj(x2−y2))∑j∈Z1¯βje−iαj(x1−z1)−i¯βj(x2−z2) ds(x)+116π2∫Γ+h∑j:βj>01βjeiαj(x1−y1)cos(βj(x2−y2))∑j∈Ze−iαj(x1−z1)−i¯βj(x2−z2) ds(x)=i16π2∑j1:βj1>0∑j2∈Z1¯βj2sin(βj1(h−y2))e−i¯βj2(h−z2)eiαj2z1−iαj1y1∫π−πei(j1−j2)x1 dx1+116π2∑j1:βj1>0∑j2∈Z1βj1cos(βj(h−y2))e−i¯βj2(h−z2)eiαj2z1−iαj1y1∫π−πei(j1−j2)x1 dx1. |
Note that
∫π−πei(j1−j2)x1 dx1={2πif j1=j2,0if j1≠j2, | (3.6) |
and therefore,
∫Γ+h∂F(x,y)∂ν¯G(x,z)−F(x,y)∂¯G(x,z)∂ν ds(x)=i8π∑j:βj>01βjsin(βj(h−y2))e−iβj(h−z2)eiαj(z1−y1)+18π∑j:βj>01βjcos(βj(h−y2))e−iβj(h−z2)eiαj(z1−y1)=18π∑j:βj>01βjeiαj(z1−y1)+iβj(z2−y2). |
Similarly, for x∈Γ−h and y∈Ωh, we have x2<y2, and we can show that
F(x,y)=18π2∑j:βj>01βjeiαj(z1−y1)cos(βj(z2−y2)). |
Moreover, by the same reasoning, we can also show that
∫Γ−h∂F(x,y)∂ν¯G(x,z)−F(x,y)∂¯G(x,z)∂ν ds(x)=18π∑j:βj>01βjeiαj(z1−y1)−iβj(z2−y2). |
Therefore,
∫Γ+h∪Γ−h∂F(x,y)∂ν¯G(x,z)−F(x,y)∂¯G(x,z)∂ν ds(x)=18π∑j:βj>0(eiαj(z1−y1)+iβj(z2−y2)+eiαj(z1−y1)−iβj(z2−y2))=14π∑j:βj>01βjeiαj(z1−y1)cos(βj(z2−y2))=F(z,y). |
The reason why (3.4) helps us define the imaging function is as follows. Using the Rayleigh expansion of the Green's function, the left-hand side of (3.4) becomes
∫Γ+h∪Γ−h∂F(x,y)∂ν¯G(x,z)−F(x,y)∂¯G(x,z)∂ν ds(x)=2π∑j:βj>0βj(¯g+j(z)g+j(y)+¯g−j(z)g−j(y)). |
Then, as done in [20], we can show that
2π∑j:βj>0βj(¯g+j(z)u+j(l)+¯g−j(z)u−j(l))=k2∫DF(z,y)q(y)u(y,l) dy. |
So, if we define the imaging function as
I(z):=N∑l=1|∑j:βj>0βj(¯g+j(z)u+j(l)+¯g−j(z)u−j(l))|p, |
we would have
I(z)=N∑l=1|k22π∫DF(z,y)q(y)u(y,l) dy|p. |
From this equation, we can expect the imaging function to exhibit behaviors similar to the kernel F(z,y). We will discuss this in detail shortly. In the case of data measured on either Γ+h or Γ−h, we will only have either u+j or u−j, respectively. A natural way to modify the imaging function is to omit the term for which data is not available. For example, when data is measured only on Γ+, the term ¯g−j(z)u−j(l) will be dropped.
For z∈Ωh, we define the (general) imaging function I(z) as
I(z):=N∑l=1|∑j:βj>0βj(χ+(t)u+j(l)¯g+j(z)+χ−(t)u−j(l)¯g−j(z))|p, | (3.7) |
where p∈N is used to sharpen the resolution of the imaging function, t∈{U, L, B} (abbreviations for Upper, Lower, and Both), and
χ+(t)={1,for t=U, B0,for t=L,χ−(t)={1,for t=L, B0,for t=U. |
The choice of t here allows the imaging function to apply to scattering data measured on either one side (Upper or Lower) or both sides (Both) of the periodic structures. The imaging function defined above satisfies the following property.
Theorem 2. For all z∈Ωh, the imaging function I(z) satisfies
I(z)=N∑l=1|k2∫DFt(z,y)q(y)u(y,l)dy|p |
where
Ft(z,y)={116π2∑j:βj>01βjeiαj(z1−y1)+iβj(z2−y2)if t=U,116π2∑j:βj>01βjeiαj(z1−y1)−iβj(z2−y2)if t=L,18π2∑j:βj>01βjeiαj(z1−y1)cos(βj(z2−y2))if t=B. | (3.8) |
Remark 3. Graphical observations display that the kernel functions (3.8) exhibit a peak when z≈y and decay rapidly as z and y are apart from each other. Moreover, this peak becomes more distinct with larger values of k, as shown in Figure 1.
The behavior of the kernel can be justified in some simple cases. For example, in the context of near-field measurements, where Γ±h are very close to periodic scatterer D, we can reasonably assume that z2≈y2. Furthermore, if we consider α≈0 in the incident field, the kernel FU(z,y) simplifies to
FU(z,y)≈116π2∑j:|j|<keij(z1−y1)√k2−j2=116π2k+18π2∑1≤j<kcos(j(z1−y1))√k2−j2. |
When z1=y1, the terms cos(j(z1−y1)) all reach their maximum value simultaneously. However, for z1≠y1, these terms also have other maxima, but they no longer align at the same point as j varies, since the periods of the cosine functions differ. Therefore, FU(z,y) will have a larger value when z≈y. As z moves farther from y, the cosine terms are more likely to differ in sign at z, leading to cancellations in their sum. As a result, F(z,y) becomes significantly smaller. A larger k results in the summation of more maximal values when z≈y and more cancellations as z and y are apart from each other. Thus, the peak of FU(z,y) appears more distinct for larger k, as illustrated in Figure 1. This justification certainly applies to Ft(z,y) for t=L or t=B. Based on the behavior of Ft(z,y), we expect I(z) to take relatively small values when z is outside D and significantly larger values when z is inside D.
It is worth noting that Theorem 2 can be proved using the integral representation (3.4) by following a similar approach as in [20]. However, in this paper, we propose an alternative simple proof. This proof does not rely on Green's identities, which allows us to study the imaging function using scattering data measured from only one side of the periodic structure.
Proof. Case t=U. Note that the Rayleigh coefficients of the scattered field satisfy
u+j(l)=12π∫Γ+husc(x,l)e−iαjx1ds(x). | (3.9) |
Substituting the Lippmann-Schwinger equation (3.1) into the integral (3.9) gives
u+j(l)=12π∫Γ+h(k2∫DGk,α(x,y)q(y)u(y,l)dy)e−iαjx1ds(x). |
Note that the integrand is bounded and D and Γ+h are bounded sets. Therefore, by Fubini's theorem,
u+j(l)=k2∫D(12π∫Γ+hGk,α(x,y)e−iαjx1 ds(x))q(y)u(y,l)dy=k2∫Dg+j(y)q(y)u(y,l)dy. |
Plugging this into the definition of I(z) gives
I(z)=N∑l=1|∑j:βj>0βj¯g+j(z)u+j(l)|p=N∑l=1|k2∫D(∑j:βj>0βj¯g+j(z)g+j(y))q(y)u(y,l)dy|p. |
By (3.3), for j such that βj>0,
¯g+j(z)g+j(y)=116π2β2jeiαjz1+iβj(z2−h)e−iαjy1−iβj(y2−h)=116π2β2jeiαj(z1−y1)+iβj(z2−y2). |
Therefore,
∑j:βj>0βj¯g+j(z)g+j(y)=116π2∑j:βj>01βjeiαj(z1−y1)+iβj(z2−y2)=FU(z,y), | (3.10) |
and thus,
I(z)=N∑l=1|k2∫DFU(z,y)q(y)u(y,l)dy|p. |
Case t=L. By similar reasoning, we can show that
∑j:βj>0βj¯g−j(z)g−j(y)=FL(z,y) | (3.11) |
and
u−j(l)=k2∫Dg−j(y)q(y)u(y,l)dy, |
and the result then follows.
Case t=B. Adding (3.10) and (3.11) together gives
∑j:βj>0βj(¯g+j(z)g+j(y)+¯g−j(z)g−j(y))=FU(z,y)+FL(z,y)=FB(z,y). |
Therefore,
I(z)=N∑l=1|∑j:βj>0βj(¯g+j(z)u+j(l)+¯g−j(z)u−j(l))|p=N∑l=1|k2∫DFB(z,y)q(y)u(y,l)dy|p. |
Theorem 2 and Remark 3 partly justify the resolution of the imaging function I(z). The next theorem shows that I(z) is stable with respect to noise in the data.
Theorem 4. For δ>0, let usc,δ(⋅,l) be the noisy scattered fields that satisfy
‖usc,δ(⋅,l)−usc(⋅,l)‖L2(Γ+h∪Γ−h)≤δ‖usc(⋅,l)‖L2(Γ+h∪Γ−h),for alll=1,2,…,N, |
and let u±j,δ(l) be the corresponding Rayleigh coefficients. We define Iδ(z) as the imaging function as in (3.7) where u±j(l) are replaced by u±j,δ(l). Then, for all z∈Ω, the following stability property holds:
|Iδ(z)−I(z)|≤Cδ,as δ→0, |
where C>0 is a constant independent of z and δ.
The proof of this theorem follows a similar approach to that in [20] and is therefore omitted here. In summary, we have introduced an imaging function for reconstructing the unknown isotropic scatterer D using scattering data measured either above or below the periodic structures. We also provided partial justification for its resolution and stability. In the next section, we extend the analysis of this imaging function to the case of anisotropic media.
In the case of scattering from anisotropic media, let Q:R2→R2×2 be the material parameter of the periodic scattering medium. We assume that Q is a matrix-valued bounded function which is 2π-periodic with respect to x1, Q is zero outside of the medium, and supp(Q) is bounded in x2. We consider the following equation:
Δusc+k2usc=−÷(Q∇u), |
where usc satisfies the radiation condition (2.3). We refer to [22] for results on the well-posedness of the direct problem under some assumption on Q and wave number k. As in the isotropic case, we assume that the direct problems is well-posed for the study of the inverse problem. Again we denote D:=supp(Q)∩Ω and the inverse problem is to determine D from given data of usc on Γ+h or Γ−h. From [24], we know that the corresponding Lippmann-Schwinger equation is given by
usc(x)=÷x∫DG(x,y)Q(y)∇u(y)dy. | (4.1) |
To simplify the presentation, we will analyze the imaging function for the case of measured data on Γ+h, and the other cases can be derived through a similar process as in the isotropic case. Recall that the imaging function for this case is
I(z):=N∑l=1|∑j:βj>0βju+j(l)¯g+j(z)|p, | (4.2) |
for z∈Ωh. The following theorem is an extension of Theorem 2 to anisotropic media.
Theorem 5. The imaging function satisfies
I(z)=N∑l=1|∫DF(z,y)⋅Q(y)∇u(y,l)dy|p, | (4.3) |
where
F(z,y):=116π2∑j:βj>01βj[αjβj]eiαj(z1−y1)+iβj(z2−y2). | (4.4) |
Proof. Plugging (4.1) into (3.9) gives
u+j(l)=12π∫Γ+h÷x(∫DG(x,y)Q(y)∇u(y,l)dy)e−iαjx1ds(x)=12π∫Γ+h÷x[∫DG(x,y)(Q11(y)∂y1u(y,l)+Q12(y)∂y2u(y,l))dy∫DG(x,y)(Q21(y)∂y1u(y,l)+Q22(y)∂y2u(y,l))dy]e−iαjx1ds(x)=12π∫Γ+h∫D∂x1G(x,y)(Q11(y)∂y1u(y,l)+Q12(y)∂y2u(y,l))dy e−iαjx1ds(x)+12π∫Γ+h∫D∂x2G(x,y)(Q21(y)∂y1u(y,l)+Q22(y)∂y2u(y,l))dy e−iαjx1ds(x). |
Swapping the integrals, we obtain,
u+j(l)=∫D(12π∫Γ+h∂x1G(x,y)e−iαjx1ds(x))(Q11(y)∂y1u(y,l)+Q12(y)∂y2u(y,l))dy+∫D(12π∫Γ+h∂x2G(x,y)e−iαjx1ds(x))(Q21(y)∂y1u(y,l)+Q22(y)∂y2u(y,l))dy. |
Since G(x,y) satisfies the radiation condition, so do ∂x1G(x,y) and ∂x2G(x,y), and their Rayleigh coefficients are iαjg+j(y) and iβjg+j(y), respectively. Thus,
u+j(l)=∫Diαjg+j(y)(Q11(y)∂y1u(y,l)+Q12(y)∂y2u(y,l))dy+∫Diβjg+j(y)(Q21(y)∂y1u(y,l)+Q22(y)∂y2u(y,l))dy=∫Dig+j(y)[αjβj]⋅Q(y)∇u(y,l)dy. |
Plugging into (4.2), we obtain
I(z)=N∑l=1|∑j:βj>0βju+j(l)¯g+j(z)|p=N∑l=1|∑j:βj>0iβj∫D¯g+j(z)g+j(y)[αjβj]⋅Q(y)∇u(y,l)dy|p=N∑l=1|∫D116π2∑j:βj>01βj[αjβj]eiαj(z1−y1)+iβj(z2−y2)⋅Q(y)∇u(y,l)dy|p=N∑l=1|∫DF(z,y)⋅Q(y)∇u(y,l)dy|p. |
This completes the proof.
The kernel F(z,y) has a similar behavior as FU(z,y) defined in (3.8). This means that, for any vector v∈R2 and z,y∈Ωh, F(z,y)⋅v has a large value when z≈y and has a much smaller value when z moves away from y. Therefore, we can expect the imaging function to have a large value when the sampling point is inside the medium and a much smaller value when the sampling point moves away from the medium. We will confirm this numerically in the next section.
In this section, we present numerical results on the performance of our imaging function. We only present results for the case of transmission data, in which the incident sources are placed below the scatterer and the receivers are placed above the scatterer. Numerical results for full-aperture data can be found in [20]. We also compare our method with the orthogonality sampling method (OSM) and the factorization method (FM). Recall that the imaging function for our method is
Iδ(z)=N∑l=1|∑j:βj>0βju+δ,j(l)¯g+j(z)|p, |
and the imaging function of the OSM is given by
IOSM(z):=N∑l=1|∫Γ+husc,δ(x,l)¯G(x,z)ds(x)|p. |
For the factorization method, we employ a singular value decomposition and regularize the method by truncating all singular values that are less than 0.1. The parameters used in the simulation are as follows.
● h=1, α=0.
● Wave number: k=10.5.
● Sampling domain: (−π,π)×(−1,1), partitioned into a grid of 128×64 sampling points.
● Number of incident sources: N=128, placed on Γ−3=(−π,π)×{−3}.
● Number of receivers: 128, placed on Γ+1=(−π,π)×{1}.
● Noise level: δ=20%. The noise is added to the scattered fields according the following model:
usc,δ=usc+δ‖usc‖L2n, |
where n:Γ+h→C is the noise function whose values are random numbers such that ‖n‖L2=1.
● Exponent of the imaging functions: p=5.
We refer to [20] for a detailed study of the behavior of the imaging function for different sets of parameters, including different values of h, α, k, and δ, as well as different numbers of incident sources and receivers. The imaging function behaves similarly in the case of limited-aperture data with respect to these parameters.
To generate synthetic data for the inverse problem, we solve the direct problem using a spectral method studied in [25]. We compute the scattered fields at the receivers' location, add noise to them, and then compute the corresponding Rayleigh coefficients via (3.9). The incident fields we used are of the form
uin(x,l)=G(x,xl),x∈Ωh, |
where xl is the receivers' location, l=1,…,N.
First, we show numerical examples of some isotropic media. Recall that in this case, the medium is characterized by the function q(x). In all examples, the function q(x) has the form
q(x)={0.5if x∈D,0if x∉D. |
We consider four periodic media with four different geometries D.
Aligned ellipses (Figure 2). This medium consists of two horizontally aligned ellipses in each period. The radii of each ellipse are π/4 and 0.4.
Aligned squares (Figure 3). This medium consists of two horizontally aligned squares in each period. The side of each square is 0.4.
Aligned crosses (Figure 4). This medium consists of two horizontally aligned crosses in each period. Each cross is made up of one horizontal 0.55×0.18 rectangle and vertical 0.18×0.45 rectangle.
Aligned kites (Figure 5). This medium consists of a kite-shaped object in each period.
In all cases, the proposed imaging function outperforms the OSM. The FM simply fails because of a high level of noise and a lack of data. Note again that we aim to solve the inverse problem under a lack of data, which is extremely challenging. Nevertheless, the proposed method is able to provide reasonable results.
In this part, we present results for some anisotropic media. In this case, the medium is characterized by a 2×2 matrix-valued function Q(x). In all examples, Q(x) has the form
Q(x)={diag(0.5,0.25)if x∈D,0if x∉D, |
where diag(0.5,0.25) is a 2×2 diagonal matrix whose first and second diagonal entries are 0.5 and 0.25, respectively. We consider the same geometries as in the isotropic case. Figures 6–9 show the reconstructions of the proposed imaging function versus those of the OSM and FM. The comparison result is similar to the isotropic case, with the proposed method outperforming both the OSM and FM.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research is partially supported by the National Science Foundation Grants DMS-2208293 and DMS-2243854.
The authors declare there are no conflicts of interest.
[1] | W. Dorfler, A. Lechleiter, M. Plum, G. Schneider, C. Wieners, Photonic Crystals: Mathematical Analysis and Numerical Approximation, Springer, 2012. https://doi.org/10.1007/978-3-0348-0113-3 |
[2] |
T. Arens, N. Grinberg, A complete factorization method for scattering by periodic structures, Computing, 75 (2005), 111–132. https://doi.org/10.1007/s00607-004-0092-0 doi: 10.1007/s00607-004-0092-0
![]() |
[3] |
G. Bao, T. Cui, P. Li, Inverse diffraction grating of Maxwell's equations in biperiodic structures, Opt. Express, 22 (2014), 4799–4816. https://doi.org/10.1364/OE.22.004799 doi: 10.1364/OE.22.004799
![]() |
[4] |
F. Cakoni, H. Haddar, T. P. Nguyen, New interior transmission problem applied to a single Floquet–Bloch mode imaging of local perturbations in periodic media, Inverse Probl., 35 (2019), 015009. https://doi.org/10.1088/1361-6420/aaecfd doi: 10.1088/1361-6420/aaecfd
![]() |
[5] |
H. Haddar, T. P. Nguyen, Sampling methods for reconstructing the geometry of a local perturbation in unknown periodic layers, Comput. Math. Appl., 74 (2017), 2831–2855. https://doi.org/10.1016/j.camwa.2017.07.015 doi: 10.1016/j.camwa.2017.07.015
![]() |
[6] |
A. Lechleiter, D. L. Nguyen, Factorization method for electromagnetic inverse scattering from biperiodic structures, SIAM J. Imag. Sci., 6 (2013), 1111–1139. https://doi.org/10.1137/120903968 doi: 10.1137/120903968
![]() |
[7] |
D. L. Nguyen, The factorization method for the Drude-Born-Fedorov model for periodic chiral structures, Inverse Probl. Imaging, 10 (2016), 519–547. https://doi.org/10.3934/ipi.2016010 doi: 10.3934/ipi.2016010
![]() |
[8] |
D. L. Nguyen, T. Truong, Imaging of bi-anisotropic periodic structures from electromagnetic near field data, J. Inverse Ill-Posed Probl., 30 (2022), 205–219. https://doi.org/10.1515/jiip-2020-0114 doi: 10.1515/jiip-2020-0114
![]() |
[9] |
T. P. Nguyen, Differential imaging of local perturbations in anisotropic periodic media, Inverse Probl., 36 (2020), 034004. https://doi.org/10.1088/1361-6420/ab2066 doi: 10.1088/1361-6420/ab2066
![]() |
[10] | K. Sandfort, The Factorization Method for Inverse Scattering from Periodic Inhomogeneous Media, Ph.D thesis, Karlsruher Institut für Technologie, 2010. https://doi.org/10.5445/KSP/1000019400 |
[11] |
J. Yang, B. Zhang, R. Zhang, A sampling method for the inverse transmission problem for periodic media, Inverse Probl., 28 (2012), 035004. https://doi.org/10.1088/0266-5611/28/3/035004 doi: 10.1088/0266-5611/28/3/035004
![]() |
[12] | A. Kirsch, N. Grinberg, The Factorization Method for Inverse Problems, Oxford University Press, 2008. https://doi.org/10.1093/acprof: oso/9780199213535.001.0001 |
[13] |
D. Colton, A. Kirsch, A simple method for solving inverse scattering problems in the resonance region, Inverse Probl., 12 (1996), 383–393. https://doi.org/10.1088/0266-5611/12/4/003 doi: 10.1088/0266-5611/12/4/003
![]() |
[14] |
A. Kirsch, Characterization of the shape of a scattering obstacle using the spectral data of the far field operator, Inverse Probl., 14 (1998), 1489–1512. https://doi.org/10.1088/0266-5611/14/6/009 doi: 10.1088/0266-5611/14/6/009
![]() |
[15] |
X. Jiang, P. Li, Inverse electromagnetic diffraction by biperiodic dielectric gratings, Inverse Probl., 33 (2017), 085004. https://doi.org/10.1088/1361-6420/aa76b9 doi: 10.1088/1361-6420/aa76b9
![]() |
[16] |
R. Griesmaier, Multi-frequency orthogonality sampling for inverse obstacle scattering problems, Inverse Probl., 27 (2011), 085005. https://doi.org/10.1088/0266-5611/27/8/085005 doi: 10.1088/0266-5611/27/8/085005
![]() |
[17] |
I. Harris, D. L. Nguyen, Orthogonality sampling method for the electromagnetic inverse scattering problem, SIAM J. Sci. Comput., 42 (2020), B722–B737. https://doi.org/10.1137/19M129783X doi: 10.1137/19M129783X
![]() |
[18] |
K. Ito, B. Jin, J. Zou, A direct sampling method to an inverse medium scattering problem, Inverse Probl., 28 (2012), 025003. https://doi.org/10.1088/0266-5611/28/2/025003 doi: 10.1088/0266-5611/28/2/025003
![]() |
[19] |
R. Potthast, A study on orthogonality sampling, Inverse Probl., 26 (2010), 074015. https://doi.org/10.1088/0266-5611/26/7/074015 doi: 10.1088/0266-5611/26/7/074015
![]() |
[20] |
D. L. Nguyen, K. Stahl, T. Truong, A new sampling indicator function for stable imaging of periodic scattering media, Inverse Probl., 39 (2023), 065013. https://doi.org/10.1088/1361-6420/acce5f doi: 10.1088/1361-6420/acce5f
![]() |
[21] |
D. L. Nguyen, T. Truong, A stable imaging functional for anisotropic periodic media in electromagnetic inverse scattering, SIAM J. Appl. Math., 84 (2024), 1631–1657. https://doi.org/10.1137/23M1577080 doi: 10.1137/23M1577080
![]() |
[22] |
A. S. Bonnet-Bendhia, F. Starling, Guided waves by electromagnetic gratings and non-uniqueness examples for the diffraction problem, Math. Methods Appl. Sci., 17 (1994), 305–338. https://doi.org/10.1002/mma.1670170502 doi: 10.1002/mma.1670170502
![]() |
[23] | D. L. Colton, R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, 2nd edition, Springer, 1998. https://doi.org/10.1007/978-3-662-03537-5 |
[24] |
A. Lechleiter, D. L. Nguyen, Volume integral equations for scattering from anisotropic diffraction gratings, Math. Methods Appl. Sci., 36 (2013), 262–274. https://doi.org/10.1002/mma.2585 doi: 10.1002/mma.2585
![]() |
[25] |
A. Lechleiter, D. L. Nguyen, A trigonometric Galerkin method for volume integral equations arising in TM grating scattering, Adv. Comput. Math., 40 (2014), 1–25. https://doi.org/10.1007/s10444-013-9295-2 doi: 10.1007/s10444-013-9295-2
![]() |