In this paper, we develop a deep learning approach for the accurate solution of challenging problems of near-field microscopy that leverages the powerful framework of physics-informed neural networks (PINNs) for the inversion of the complex optical parameters of nanostructured environments. Specifically, we show that PINNs can be flexibly designed based on full-vector Maxwell’s equations to inversely retrieve the spatial distributions of the complex electric permittivity and magnetic permeability of unknown scattering objects in the resonance regime from near-field data. Moreover, we demonstrate that PINNs achieve excellent convergence to the true material parameters under both plane wave and point source (localized) excitations, enabling parameter retrieval in scanning near-field optical microscopy. Our method is computationally efficient compared to traditional data-driven deep learning approaches as it requires only a single dataset for training. Furthermore, we develop and successfully demonstrate adaptive PINNs with trainable loss weights that largely improve the accuracy of the inverse reconstruction for high-index materials compared to standard PINNs. Finally, we demonstrate the full potential of our approach by retrieving the space-dependent permittivity of a three-dimensional unknown object from near-field data. The presented framework paves the way to the development of a computationally driven, accurate, and non-invasive platform for the simultaneous retrieval of the electric and magnetic parameters of resonant nanostructures from measured optical images, with applications to biomedical imaging, optical remote sensing, and characterization of metamaterial devices.

In the past few decades, the engineering of electromagnetic waves in optical materials heavily relied on either analytical theories or numerical methods to obtain the solutions of physics-based partial differential equation (PDE) models. However, it has become increasingly difficult to apply these traditional approaches to complex optical structures and heterogeneous media, particularly in relation to the challenging inverse problems of near-field optical microscopy with many applications to biomedical imaging, material characterization, and nano-optical device inspection.1–4 Specifically, the parameter retrieval problem of near-field microscopy consists in estimating the properties of a scattering object, usually identified by its shape and dielectric permittivity, from a limited set of near-field data under different excitation conditions. Thanks to the impressive developments of the scanning near-field optical microscopy (SNOM) technique,5 it is currently possible to image the amplitude and phase of the scattered fields of complex nanostructures in the near-field zone with nanoscale spatial accuracy.2–10 In particular, by exciting and detecting photonic structures locally, near-field techniques provide a “non-invasive” approach for the characterization of complex nanostructures. However, due to the strong multiple scattering of radiation in multi-particle systems at large refractive index contrast,11 the problem of retrieving the material optical constants from near-field data becomes intrinsically non-linear and ill-posed. As a result, iterative optimization methods are proposed to solve these inverse problems with good accuracy, but they are usually computationally expensive.12–18 On the other hand, there is a growing interest in developing deep learning (DL) algorithms for electromagnetic wave engineering. This rapidly emerging approach includes training artificial neural networks (ANNs) to solve the photonics inverse problems.19–29 Although demonstrated successfully at solving several inverse problems,23,24 DL methods are essentially data-driven techniques that require a time-consuming training process in order to instruct ANNs using massive datasets. In order to improve on purely data-driven DL methods, it is important to constrain them by leveraging the underlying physics of the problems, thus relaxing the burden on the training and data acquisition steps. Therefore, it is critical to build a robust framework that efficiently integrates powerful ANN architectures with the physical laws that fundamentally constrain the complex parameter retrieval problems of near-field microscopy. In this context, the physics-informed neural network (PINN) is a general framework developed for solving both forward and inverse problems that are mathematically modeled by arbitrary PDEs of integer or fractional orders.30–32 In our previous work, we have applied PINNs to retrieve purely real permittivities of lossless materials from real-valued electric field observations under plane wave excitation.33 

In this paper, we develop a more general PINN framework for solving the parameter retrieval problems of near-field optical microscopy that are of direct interest to biomedical imaging and nanotechnology. In particular, we address and demonstrate the accurate retrieval of the complex electric permittivity ɛr of resonant nanostructures based on complex wave equations using complex electric field (synthetic) data obtained from forward finite element method (FEM) simulations. Furthermore, we demonstrate the retrieval of the complex magnetic permeability μr of resonant nanostructures based on the full inversion of Maxwell’s equations. We show that PINNs retrieve correctly the material parameters from near-field data sampled under both plane wave and localized (line current source) excitations, which can enable inverse parameter retrieval using the dual-SNOM technique.2,8,10 Importantly, we show the simultaneous inverse retrieval of the space-dependent material parameters ɛr and μr without prior shape information in a two-dimensional (2D) geometry. In this context, an adaptive PINN algorithm is proposed and developed to improve the stability and accuracy in retrieving high-index material parameters in a regime where the standard PINN method fails. Finally, we demonstrate the full potential of our method by successfully retrieving the complex permittivity profile of an unknown three-dimensional (3D) scattering object from sampled synthetic data. We remark that the framework shown here is more general than the one shown in our previous work33 and can be used to simultaneously retrieve the space-dependent complex optical parameters of unknown objects with electric and magnetic resonant responses in the presence of losses and under different excitation conditions. Here, we demonstrate our approach by retrieving the optical constants for linear and isotropic materials. However, the developed framework can be naturally extended to retrieve nonlinear optical and anisotropic parameters by implementing more general wave equations.34 All the implementations of the developed PINN algorithms used in this paper are obtained within the powerful DeepXDE library.31 

The general PINN algorithm utilized for solving inverse PDE problems is schematically illustrated in Fig. 1. We first construct an artificial neural network (ANN) with input vector x = (x1, y1x2, y2; …; xN, yN) whose coordinates denote either points on a grid or randomly distributed points over the investigated domain Ω and output û(x;θ,ξ) that is a surrogate of the PDE solution u(x).31 Here, θ denotes a vector containing all the weights and biases in the ANN and ξ describes the unknown parameters in the PDEs that need to be retrieved. A simple fully connected neural network (FCNN) is employed here, but the method can conveniently be extended to accommodate more complex ANN architectures. Crucially, we have built the loss function that constrains û to satisfy the PDEs describing the electromagnetic physics of the considered problems.

FIG. 1.

(a) Schematics of the PINN solving the parameter retrieval problem for near-field microscopy. The FCNN (x, yθξ) denotes the fully connected neural network with its output as the PDE surrogate solution for the inverse problem. The “Loss” symbol in the bottom represents the loss function that restricts û to satisfy the PDEs, complex field data, and boundary conditions (BCs). During the training process, both θ and unknown object material parameters ξ in the neural networks are optimized until the value of the loss function is below certain threshold σ.

FIG. 1.

(a) Schematics of the PINN solving the parameter retrieval problem for near-field microscopy. The FCNN (x, yθξ) denotes the fully connected neural network with its output as the PDE surrogate solution for the inverse problem. The “Loss” symbol in the bottom represents the loss function that restricts û to satisfy the PDEs, complex field data, and boundary conditions (BCs). During the training process, both θ and unknown object material parameters ξ in the neural networks are optimized until the value of the loss function is below certain threshold σ.

Close modal

As a first example, we consider the near-field microscopy parameter retrieval problem of retrieving the complex relative permittivity ɛr defined on a domain Ω ⊂ R2 with the constraint as follows:

(1)

with boundary conditions B(u,x)=0 on Ω, where ɛr corresponds to the ξ parameter in PINNs and the function f is derived from wave equation as it will be discussed in Sec. II A 1. The derivatives of û(x;θ,εr) in Eq. (1) are obtained using the auto differentiation of the ANN, which is already implemented in the TensorFlow package.35 We define the loss function that constraints our PINN formulations by

(2)

where wf, wb, and wi are the loss weights and

(3)
(4)
(5)

where Lf, Li, and Lb denote the L2 norm of residuals for the PDEs, complex field observations, and BCs, respectively. The symbols Re{·} and Im{·} denote the real and imaginary parts of a complex quantity, respectively. We can obtain the complex field observations uobs(x) from experimental or numerical simulations (i.e., synthetic data). The quantities Tf, Ti, and Tb denote the residual points for Lf, Li, and Lb, respectively.31 In the last step, we train the neural networks of PINNs to search for the parameters θ and ɛr that minimize the total loss function specified in Eq. (2). As we will show in Sec. II, the proposed framework solves the parameter retrieval problem of near-field microscopy using only one set of complex field observations. We remark that using PINNs, we can solve a complex inverse problem by adding a small computational cost compared to the solution of the associated forward one30,31,36 since the only difference between the two is the introduction of the extra loss term Li.

1. Retrieval of the complex electric permittivity

We recently utilized the powerful PINN method for solving the inverse Mie scattering problem limited to lossless materials.33 Here, we extend the approach by taking into account the losses of the materials, which require the more difficult inversion of complex optical parameters. We start by considering the case of retrieving the complex permittivity of a single dielectric cylinder with a diameter comparable with the wavelength of light. Specifically, we study a cylinder with radius r = 2 µm and ɛr = 3 + j(1) under TM polarized plane wave excitation at wavelength λ = 3 µm. We obtain synthetic data by performing forward simulations using the FEM modeling33 (see Sec. IV for additional details on the FEM simulations). The ɛr real part profile of the simulated structure is shown in Fig. 2(a). We denote the region of vacuum (Re{ɛr} = 1) as Ω1 and the region occupied by the cylinder (Re{ɛr} = 3) as Ω2.

FIG. 2.

(a) Real part of the ɛr profile used for the FEM forward scattering simulation. (b) and (c) Real and imaginary parts of the complex electric field Ez, respectively, used to train PINNs. The blank regions in (b) and (c) indicate that the Ez data inside the cylinder are excluded. (d) Inverse retrieval of the complex dielectric function with respect to the number of iterations. (e) and (f) Real and imaginary parts of the complex Ez field reconstructed by PINNs after 104 iterations.

FIG. 2.

(a) Real part of the ɛr profile used for the FEM forward scattering simulation. (b) and (c) Real and imaginary parts of the complex electric field Ez, respectively, used to train PINNs. The blank regions in (b) and (c) indicate that the Ez data inside the cylinder are excluded. (d) Inverse retrieval of the complex dielectric function with respect to the number of iterations. (e) and (f) Real and imaginary parts of the complex Ez field reconstructed by PINNs after 104 iterations.

Close modal

Dynamic Maxwell’s equations allow one to derive the wave equation for a non-homogeneous medium in the following form:37 

(6)

where E is the electric field, H is the magnetic field, and ɛ(r) = ɛ0ɛr(r) and μ(r) = μ0μr(r) are the space-dependent medium permittivity and permeability, respectively. The symbols μ0 and ɛ0 denote the permeability and permittivity of free space, respectively. Without loss of generality, we first study the parameter retrieval in 2D geometries for retrieving the complex parameter ɛr under the TM polarization excitation E(r) = Ez(x, y), which yields E · ∇ln ɛ(x, y) = 0. In addition, we assume, at first, that a non-magnetic object with relative permeability μr(r) = 1 for which ∇ln μ = 0 yields the following wave equation:

(7)

Separating the real and imaginary parts of the wave equation, we obtain the following PDE model that we have implemented in our PINNs:

(8)
(9)

where k0=2πλ is the incident wave number. Since the shape of the object is known a priori, we denote by ɛr1 and ɛr2 the complex homogeneous permittivity in regions Ω1 and Ω2, respectively, and we impose the electromagnetic BCs at the boundary Ω as follows:

(10)
(11)
(12)
(13)

where Ez(k),(k=1,2) are the complex electric fields in domain Ωk and r is the radial component in polar coordinates with its origin at the center of the cylinder. The real and imaginary parts of Ez obtained from FEM simulations and utilized for training PINNs are displayed in Figs. 2(b) and 2(c), respectively. We sample the complex Ez on a square grid in the Ω1 region with resolution Δx = 0.02λ as the training dataset, which is achievable using current near-field microscopy techniques.22,23 We have also investigated the robustness of the retrieval with respect to different values of the spatial resolution Δx in Sec. 2.3 of the supplementary material. We set the initial value ɛr2 = 1 + j(−1) and train the FCNN by minimizing the loss function specified in Eq. (2). Further details on the training methods and the utilized hyperparameters of the FCNN are described in Sec. IV. The retrieved values of the complex permittivity ɛr2 with respect to the iteration number are shown in Fig. 2(d) where the values of the true solution are indicated by the dashed lines. The results displayed in Fig. 2 demonstrate the accurate retrieval of the complex permittivity of non-ideal lossy dielectric materials using PINNs. We show that PINNs can still accurately retrieve permittivity data starting from different initial values, as discussed in Sec. 2.2 of the supplementary material. Moreover, the developed PINN model successfully reconstructed the complex Ez distribution within Ω2 after training, as we show in Figs. 2(e) and 2(f) for the real and imaginary parts of Ez, respectively. The L2 errors of Ez between data obtained from FEM simulations and the predictions of PINNs are 104 (the L2 error definition and its exact value for Ez real and imaginary parts are discussed in Sec. 2.7 of the supplementary material). Therefore, we conclude that PINNs can be reliably extended to obtain the electric permittivity of resonant nanostructures from complex electric field data. In Sec. II A 2, we introduce an even more general PINN model for the simultaneous retrieval of the complex optical parameters of electric and magnetic scattering media.

2. Retrieval of permittivity and permeability

In this section, we consider the retrieval of both the μr and ɛr values of magnetic materials that have important applications in biomedical, environment treatment, and nanotechnology.38–41 Since the parameters μr and ɛr are coupled in Eq. (7), we now consider a PDE model based on dynamic Maxwell’s equations, where μr and ɛr appear separately. In particular, for the TM polarization, the PDEs are as follows:

(14)
(15)
(16)

We denote the relative permeability and permittivity of region Ωk as μrk and ɛrk (k = 1, 2), respectively. Here, Hx and Hy represent the x and y components of the magnetic field, respectively. We simulate a cylinder with the same geometry as shown in Fig. 2(a), and we set ɛr1 = μr1 = 1, ɛr2 = 1, and μr2 = 2.5 + j(0.6). The same notations of Ω1 for the vacuum region and Ω2 for the cylinder region are used. We can directly apply the following BCs at the cylinder boundary Ω2 to constrain the PINN’s retrieval of the optical constants using

(17)
(18)
(19)
(20)

where n is the unit vector normal to Ω and Hk=(Hx(k),Hy(k)),(k=1,2) is the magnetic field vector in the region Ωk.

We perform the FEM simulations with details specified in Sec. IV and obtain the complex Ez, Hx, and Hy field data for training the network. Similar to the approach of retrieving ɛr, we sampled these fields only in the region Ω2 considered as the training dataset. We train the FCNN (see Sec. IV for more details) to retrieve the complex μr2 by fixing ɛr1 = μr1 = 1 and ɛr2 = 1. μr2 starts from the initial value μr2 = 2. The reconstructed complex Ez profile obtained from the PINNs is shown in Sec. 1.1 of the supplementary material. We display the reconstructed complex Hx and Hy field profiles in Figs. 3(a) and 3(b) and Figs. 3(c) and 3(d), respectively. We show the parameter retrieval during the training process in Fig. 3(e), where very good convergence to the complex μr2 true solutions (indicated by the dashed lines) is obtained. Moreover, we demonstrate that the retrieval results also converge to the true solutions for different initial values in Sec. 2.2 of the supplementary material. We quantify the convergence by computing the maximum L2 error norm between the reconstructed fields from PINNs and FEM simulations, which we found to be 8 × 10−3 (detailed values are reported in Sec. 2.7 of the supplementary material). The scaling of the total loss with respect to the iteration numbers is displayed in Fig. 3(f) where a satisfactory loss of ≈10−2 is reached. Moreover, in Sec. 2.4 of our supplementary material, we show that PINNs can retrieve simultaneously μr and ɛr for a cylinder with a training dataset sampled over both Ω1 and Ω2. Therefore, we have shown that the proposed framework can be generally implemented to solve the near-field parameter retrieval problem for both the electric and magnetic properties of an object of a given shape based on Maxwell’s equations under plane wave excitation. In Sec. II A 3, we consider how to solve such a problem under a local source excitation, which is directly relevant to the applications of the SNOM technique for inverse parameter retrieval.

FIG. 3.

(a) and (b) Real and imaginary parts of the complex magnetic field Hx component used in PINNs, respectively. (c) and (d) Real and imaginary parts of the complex magnetic field Hy component used in PINNs, respectively. (e) Retrieval of the complex permeability μr2 with respect to the number of iterations. The dashed lines indicate the true values of retrieved parameters. (f) Total loss value during the training process with respect to the iteration number.

FIG. 3.

(a) and (b) Real and imaginary parts of the complex magnetic field Hx component used in PINNs, respectively. (c) and (d) Real and imaginary parts of the complex magnetic field Hy component used in PINNs, respectively. (e) Retrieval of the complex permeability μr2 with respect to the number of iterations. The dashed lines indicate the true values of retrieved parameters. (f) Total loss value during the training process with respect to the iteration number.

Close modal

3. Permittivity retrieval using the dual-SNOM setup

In Sec. II A 2, we discuss the parameter retrieval of complex ɛr and μr from the near-field data under plane wave excitation. However, in many applications of the SNOM technique, complex nanostructures are often illuminated using an excitation probe in the near-field.2–10 In particular, we address here the parameter retrieval problem of the dual-SNOM technique2,8,10 that uses two localized probes for the simultaneous excitation and detection of the investigated nanostructures. The FEM setup for simulating the parameter retrieval using the dual-SNOM is schematically illustrated in Fig. 4(a). We illuminate the same photonic structure as in Sec. II A 1 with a line current source along the z-axis (out-of-plane) at the position indicated by the small red dot that emulates the excitation probe. The near-field information for training PINNs is then collected by the detecting probe that is scanned across the detection area (white dashed square, excluding the cylinder domain). We keep the same field sampling resolution as in the previous examples. The wave equation models in Eqs. (8) and (9) can still be implemented because the electric field is still TM polarized E(r) = Ez(x, y). Furthermore, the BCs described by Eqs. (10)(13) can also be applied. We train an FCNN with the same hyperparameters and training method used in Sec. II A 1. The obtained complex electric field Ez profiles from the PINN are shown in Figs. 4(b) and 4(c), where we retrieve the Ez profiles also inside the cylinder region (based uniquely on external field data). We evaluate the L2 errors for the Ez complex field between the FEM simulation and PINNs to be 1×104 (detailed values are given in Sec. 2.7 of the supplementary material). The complex ɛr retrieval with respect to the number of iterations is shown in Fig. 4(d), demonstrating the rapid convergence to the true values using the dual-SNOM configuration. The PINNs based on the wave equation can hence successfully retrieve ɛr using only the external near-field data under different excitation conditions. Furthermore, the simultaneous retrieval of ɛr and μr for photonic structures can also be achieved following the same approach. Therefore, the PINN framework can be directly applied to solve the electric and magnetic material parameter retrieval problems under the dual-SNOM setup.

FIG. 4.

(a) Schematic of FEM simulation for the dual-SNOM setup. (b) and (c) Real and imaginary parts of the complex electric field Ez in the detection area obtained from PINNs, respectively. (d) Retrieval of the complex permittivity ɛr2 with respect to the number of iterations. The dashed lines indicate the true values of retrieved parameters.

FIG. 4.

(a) Schematic of FEM simulation for the dual-SNOM setup. (b) and (c) Real and imaginary parts of the complex electric field Ez in the detection area obtained from PINNs, respectively. (d) Retrieval of the complex permittivity ɛr2 with respect to the number of iterations. The dashed lines indicate the true values of retrieved parameters.

Close modal

4. Parameter retrieval in the presence of noise

In this section, we study the robustness of the parameter retrieval problem in the presence of noise. Although we assumed the accurate near-field measurement (synthetic field data from FEM simulations) in Secs. II A 1II A 3, real experimental data acquisitions are subjected to uncertainties and errors in any practical measurement systems. It has been recently shown that PINNs can effectively de-noise the noisy data and obtain convergence to the ground truth for realistic magnetic resonance images36 and surface acoustic data.42 Furthermore, the Gaussian process smoothed PINNs are shown to recover the performance of the PINN for solving the Schrödinger equation with noisy and corrupted initial data.43 

Here, we show that PINNs are robust to the presence of noise in the near-field measurements for the parameter retrieval problem of electromagnetic scattering systems. We consider the presence of additive Gaussian noise in the complex Ez external field data used for the complex ɛr retrieval. Specifically, we superimpose the zero-mean Gaussian noise with standard variation σ = 0.05 (10% of the maximum Ez amplitude) to the real and imaginary parts of the complex training dataset used in Sec. II A 3. The real part of the obtained complex training dataset is shown in Fig. 5(a), and its imaginary part is displayed in Fig. 5(b). We use the same neural network as implemented for Fig. 4, while we train it using the noisy training dataset. We show the parameter retrieval process with respect to the iteration number in Fig. 5(c), where we demonstrate that the complex ɛr2 still converges to the ground truth. Moreover, the PINN removed the noise in the complex Ez data in its reconstructed fields as shown in Figs. 5(d) and 5(e). We further evaluate the L2 errors between Ez reconstructed from PINNs and the one without noise and find these values to be 0.28% and 0.23% for the real and imaginary parts, respectively. However, we obtain a total loss value of 0.5 after 1 × 104 iterations, which is larger than the one without noise. This is because we used a noisy dataset for training the PINN and reconstructed the Ez profiles without the noise. Therefore, the computed loss term Li is expected to increase. Figure 5(f) shows how the relative error between the retrieved complex ɛr2 value and its true value varies with respect to σ of the Gaussian noise. We can observe that the relative error for PINNs’ retrieval increases significantly with σ. Remarkably, we show that the PINN is capable of retrieving the optical parameters with relative error below 10% from noisy data with σ equal to 40% of the maximum Ez amplitude, yielding physically consistent reconstructions. In Sec. II B, we address the PINN approach to the complex parameter retrieval of resonant dielectric objects of unknown shapes.

FIG. 5.

(a) and (b) Real and imaginary parts of the Ez profile, respectively, superimposed with Gaussian noise. (c) Inverse retrieval of the complex dielectric function with respect to the number of iterations. (d) and (e) Real and imaginary parts of the complex Ez field reconstructed by PINNs after 104 iterations. (f) Relative error for the complex ɛr2 retrieval results with respect to the σ value of Gaussian noise. The error bars quantify the effect of different noise realizations.

FIG. 5.

(a) and (b) Real and imaginary parts of the Ez profile, respectively, superimposed with Gaussian noise. (c) Inverse retrieval of the complex dielectric function with respect to the number of iterations. (d) and (e) Real and imaginary parts of the complex Ez field reconstructed by PINNs after 104 iterations. (f) Relative error for the complex ɛr2 retrieval results with respect to the σ value of Gaussian noise. The error bars quantify the effect of different noise realizations.

Close modal

1. Retrieval of the complex permittivity profile

In this section, we address the challenging problem of retrieving not only the complex permittivity values in the resonant regime but also the unknown shapes of the scattering objects. In particular, we will show that the proposed PINN framework can be developed to retrieve the complex ɛr(x, y) profiles of unknown objects with dimensions comparable to or larger than the wavelength of the incident light by inverting the wave equations without imposing any BCs. The efficient solution of the inverse scattering problems in non-homogeneous media shown here simultaneously addresses both the parameter retrieval and the imaging problem of microscopy that are computationally prohibitive using the traditional retrieval methods.44,45 In order to demonstrate this capability using PINNs, we consider as a representative example the permittivity profile corresponding to the dimer configuration shown in Fig. 6(a). We set the larger cylinder with radius r1 = 2 µm and permittivity ɛr1 = 3 + j(1) centered at (xc1, yc1) = (0, 1 µm) and the smaller cylinder with radius r2 = 0.8 µm and permittivity ɛr2 = 6 + j(2) centered at (xc2, yc2) = (0, −2 µm). The complex field data Ez obtained from FEM simulation under TM plane wave excitation with λ = 3 µm are shown in Figs. 6(b) and 6(c) for the real and imaginary parts, respectively. More details on the FEM simulations are provided in Sec. IV.

FIG. 6.

(a) Real part of the ɛr profile for the two asymmetric dielectric dimer used for the FEM forward scattering simulation. (b) and (c) Real and imaginary parts of the complex electric field Ez obtained from FEM simulation. The wavelength of the incident plane wave is λ = 3 µm. (d) and (e) Real and imaginary parts of the retrieved permittivity profile ɛr by PINNs after 15 × 104 iterations. (f) Total loss value during the training process with respect to the iteration number in PINNs.

FIG. 6.

(a) Real part of the ɛr profile for the two asymmetric dielectric dimer used for the FEM forward scattering simulation. (b) and (c) Real and imaginary parts of the complex electric field Ez obtained from FEM simulation. The wavelength of the incident plane wave is λ = 3 µm. (d) and (e) Real and imaginary parts of the retrieved permittivity profile ɛr by PINNs after 15 × 104 iterations. (f) Total loss value during the training process with respect to the iteration number in PINNs.

Close modal

We implement the PDE model described by Eqs. (8) and (9) with a space-dependent complex relative permittivity εrx,y over the entire square domain Ω. Since the shape of the object is not known a priori, no BCs can be implemented in this case and the problem directly deals with the inversion of a non-homogeneous extended medium. We then train PINNs with only the PDE and field observation constraints and retrieve the complex εrx,y profile after the training process. Here, the complex εrx,y functions are directly the outputs of the ANN rather than trainable variables as we used in the previous examples. To retrieve εrx,y shown in Fig. 6(a), we employ the same FCNN architecture as in the last example and train it using the Adam optimizer for 1.5 × 104 iterations until the total loss drops below 10−2. The real and imaginary parts of the retrieved ɛr profile after training are shown in Figs. 6(d) and 6(e), respectively. We observe that the proposed framework retrieved successfully the shape information of the dimer setup in Fig. 6(d). Furthermore, we characterize the accuracy of the retrieved profiles by evaluating the complex ɛr inside each cylinder domain, and we obtain ɛr1 = (2.92 ± 0.27) + j(0.96 ± 0.12) and ɛr2 = (5.97 ± 0.34) + j(2 ± 0.15), which are in very good agreement with the input data. The ɛr errors estimated here are the standard deviations of the corresponding quantities within each cylinder domain. We also evaluate the L2 error between Ez obtained from PINNs and from the FEM simulations, which is 10×104 (see Sec. 2.7 of the supplementary material for further details). We display the total loss with respect to the iteration in Fig. 6(f). The rapid spikes displayed by the total loss curve during the training process visibly demonstrate the highly non-linear nature of the parameter retrieval problem for near-field microscopy. Notice that we successfully retrieved the space profile of the complex permittivity ɛr at almost no additional computational cost compared to the previous example shown in Sec. II A 1, except that here we used the complex Ez data over the entire Ω domain as our dataset.

We further demonstrate the capability of retrieving shape information for more complex geometries and near-field data. We choose a non-canonical polygon geometry as shown in Fig. 7 with a high aspect ratio of 2.6 (the ratio between two different side lengths) and sharp edges. This polygon shape represents a generally complex geometry with the training datasets containing sharp field variations compared to the dimer geometry. We set the permittivity of the polygon equal to ɛr = 6 + j(3). The complex ɛr profile of the polygon real and imaginary parts is shown in Figs. 7(a) and 7(b), respectively. We use the same training setup as implemented for Fig. 6. The real and imaginary parts of Ez used for training are shown in Figs. 7(c) and 7(f), respectively. The PINN correctly retrieves the real part of ɛr as shown in Fig. 7(d) and the imaginary part as displayed in Fig. 7(e). We further characterize ɛr inside the polygon region and obtain ɛr = (5.8 ± 0.6) + j(2.8 ± 0.3), which is in agreement with the ground truth. For better visualization of the retrieval results, we further apply a threshold constraint to the obtained complex permittivity profiles. The results are displayed in Sec. 2.5 of the supplementary material. Therefore, the developed PINN inversion models demonstrate the accurate and efficient retrieval of both the complex permittivity values and the space distributions (i.e., shape information) of scattering objects. This achievement naturally augments near-field microscopy techniques by providing a robust, computationally driven platform for solving the imaging and the parameter retrieval problem of dielectric structures simultaneously.

FIG. 7.

(a) and (b) Real and imaginary parts of the ɛr profile used for the FEM forward scattering simulation, respectively. (c) and (f) Real and imaginary parts of the complex electric field Ez used to train PINNs, respectively. (d) and (e) Real and imaginary parts of the retrieved permittivity profile by PINNs after 15 × 104 iterations, respectively.

FIG. 7.

(a) and (b) Real and imaginary parts of the ɛr profile used for the FEM forward scattering simulation, respectively. (c) and (f) Real and imaginary parts of the complex electric field Ez used to train PINNs, respectively. (d) and (e) Real and imaginary parts of the retrieved permittivity profile by PINNs after 15 × 104 iterations, respectively.

Close modal

2. Simultaneous retrieval of permittivity and permeability profiles

In this section, we demonstrate how to improve the previous PINN setup in order to retrieve simultaneously both the ɛr(x, y) and μr(x, y) spatial profiles, providing both electric and magnetic optical parameters together with shape information for applications to inverse near-field microscopy. In this case, we must implement Eqs. (14)(16) and retrieve the space-dependent functions εrx,y and μrx,y defined over the entire domain Ω. Since we are dealing with a full-domain, non-homogeneous retrieval problem, no BCs need to be applied here.

The investigated dimer has the same dimensions as previously shown in Fig. 6(a) except that here we set the optical constants of the two cylinders as ɛr1 = 1, μr1 = 1.5 + j(0.5) and ɛr2 = 6 + j(3), μr2 = 1, where one is purely magnetic, while the other one is purely dielectric. We run the FEM simulations with settings detailed in Sec. IV. The FEM simulation results for the real and imaginary components of Hx used for training PINNs are shown in Figs. 8(a) and 8(b), respectively. We display the training datasets Hy of real and imaginary parts in Figs. 8(c) and 8(d), respectively. The complex Ez field data used for training are shown in Sec. 1.2 of the supplementary material. The FCNN parameters and training details are given in Sec. IV. We trained the FCNN for 1.5 × 104 iterations before reaching a satisfactory total loss value of 1.5 × 10−2. The real part of permittivity and permeability spatial profiles retrieved by PINNs is shown in Figs. 8(e) and 8(f), respectively, which demonstrate the accurate reconstruct of each cylinder’s shape. We show the retrieved permittivity and permeability imaginary parts in Sec. 2.1 of the supplementary material. The obtained complex ɛr(x, y) and μr(x, y) profiles inside the two cylinder domains have constant values equal to ɛr1 = (1.00 ± 0.19) + j(0 ± 0.01), μr1 = (1.48 ± 0.07) + j(0.48 ± 0.07) and ɛr2 = (5.75 ± 0.61) + j(3 ± 0.36), μr2 = (1 ± 0.04) + j(0 ± 0.05), respectively. The maximum L2 error between fields from PINN and FEM simulations is evaluated to be 8 × 10−3. Therefore, we successfully demonstrated the full retrieval of both the shape of the particle and values of the electric and magnetic parameters from synthetic electric and magnetic field data. However, in order to obtain stable results with better accuracy at large refractive index contrasts, we need to further generalize the PINN framework by introducing adaptive weights, as discussed in Sec. II C.

FIG. 8.

(a) and (b) Real and imaginary parts of the complex magnetic field Hx component, respectively. (c) and (d) Real and imaginary parts of the complex magnetic field Hy component, respectively. (e) and (f) Real part of the retrieved complex ɛr and μr profiles by PINNs after 1.5 × 104 iterations.

FIG. 8.

(a) and (b) Real and imaginary parts of the complex magnetic field Hx component, respectively. (c) and (d) Real and imaginary parts of the complex magnetic field Hy component, respectively. (e) and (f) Real part of the retrieved complex ɛr and μr profiles by PINNs after 1.5 × 104 iterations.

Close modal

In Secs. II A and B, we showed that PINNs are suitable for the retrieval of the complex ɛr and μr of resonant nanostructures from near-field observations outside the objects. However, the solutions of such complex inverse problems become progressively more inaccurate by increasing the refractive index contrast. For instance, we have shown in Sec. 2.6 of the supplementary material that the standard PINN approach loses its accuracy when increasing the ɛr and μr values of the object above a certain threshold value. Therefore, a more accurate and flexible PINN approach needs to be developed where the loss weights in Eq. (2) are not fixed but can be adaptively modified for the solution of high-index problems. In this section, we, in fact, demonstrate that additionally training the PINNs’ loss weights significantly improves the accuracy of the parameter retrieval for high-index scattering objects. It has been recently demonstrated that adaptive PINN methods can outperform standard PINNs in accurately solving PDEs with solutions containing sharp transition and sudden fronts, such as the situations encountered in phase-field PDEs.46–48 The basic idea behind adaptive PINNs is to increase the loss weights for the loss terms that are high. In particular, we apply the following updates for the loss weights at the kth time of n iterations in addition to the standard PINNs:

(21)
(22)
(23)

where η is the learning rate for the loss weights. We choose the cylinder with the same geometry as in Sec. II A 1 with ɛr = 5 + j(1) and show in Sec. 2.6 of the supplementary material that the standard PINNs fail to retrieve.

The complex ɛr retrieval results with respect to the iteration number by using the standard PINNs (ɛr2,na) and adaptive PINNs (ɛr2,a) are compared in Fig. 9(a). The same FEM simulation and normal PINN setup as in Sec. II A 1 are used. For the adaptive PINNs, we choose η = 5 and update the loss weights by every 5000 iterations (n = 5000). We use the fixed loss weight values in standard PINNs as the initial loss weight values for the adaptive PINNs. Further training details are specified in Sec. IV. We observe that at the beginning of the training process, ɛr2,a and ɛr2,na are close because the initial loss weights for the adaptive PINNs are the same as the fixed loss weights for the standard PINNs. However, as the simulation progresses further, ɛr2,a converges to its correct value, and this value is very different from ɛr2,na at the end of the simulation due to the importance of the loss weight updates. We show the reconstructed complex Ez real and imaginary profiles in Figs. 9(b) and 9(c) by using adaptive PINNs, respectively. The L2 errors with respect to FEM solutions of the Ez profiles obtained from PINNs are now as low as 1 × 10−4 and 2 × 10−4 for the real and imaginary parts, respectively. Therefore, the developed adaptive-PINN formulation is suitable for the study of the complex near-field profile of high-index scatterers and correctly retrieves their complex optical constants in situations where the standard PINNs loses their accuracy entirely. Furthermore, instead of applying the fixed loss weights with values determined using the trial and error procedure, the adaptive PINN method can balance the interplay between different loss terms automatically. We demonstrated parameter retrieval for a high-index material by using the adaptive PINNs to improve the retrieval accuracy in a 2D configuration. In Sec. IV of our paper, we introduce the implementation of the general PINN model for complex parameter retrieval of 3D objects with unknown shapes.

FIG. 9.

(a) Complex ɛr retrieval with respect to the iteration number by the normal PINN and adaptive PINNs. (b) and (c) Reconstructed real and imaginary Ez profiles obtained by adaptive PINNs, respectively.

FIG. 9.

(a) Complex ɛr retrieval with respect to the iteration number by the normal PINN and adaptive PINNs. (b) and (c) Reconstructed real and imaginary Ez profiles obtained by adaptive PINNs, respectively.

Close modal

We finally extend the PINN framework to retrieve the complex permittivity of 3D objects with unknown shapes and composition, which directly addresses the inverse retrieval of optical parameters of nanostructures used in practical biomedical and nanotechnology applications. As a representative example, we show the 3D retrieval of the permittivity profile of non-magnetic objects. However, the presented framework can be modified as shown in Sec. II AII C to additionally retrieve the complex ɛr and μr 3D profiles simultaneously. By training with the complex electric field data in 3D space and restricting the search based on the wave equation for 3D non-homogeneous media, we demonstrate that PINNs can successfully retrieve the complex permittivity of 3D objects. The implemented non-homogeneous wave equations for non-magnetic objects can be derived from Eq. (6) as

(24)

We consider here a 3D sphere with radius r1 = 2 µm and constant relative permittivity ɛr1 = 3. The complex electric field data E = (Ex, Ey, Ez) obtained from 3D FEM simulations under plane wave illumination with wavelength λ = 3 µm are used as the complex field observations to train PINNs. We display the 3D electric field real part profiles for training PINNs in Figs. 10(a)10(c). We employ a FCNN and train it over 120 000 iterations until the total loss is below 10−1. Additional details on the 3D FEM simulations and the network parameters are discussed in our Sec. IV. The retrieved 3D permittivity is shown in Fig. 10(d). The obtained non-homogeneous permittivity inside the 3D sphere region is ɛr = 2.57 ± 0.45, which is in qualitatively good agreement with the ground truth value ɛ1. This result, which is only limited by our available computational power (we intentionally used a desktop computer for conducting this work as specified in Sec. IV), can be further improved by increasing the sampling points for the electric fields in 3D. We conclude by remarking that a similar approach can be applied to extend this PINN framework for the simultaneous 3D retrieval of ɛr and μr based on 3D near-field microscopy data.

FIG. 10.

(a)–(c) Planar cross sections of the 3D electric field distribution (real part) profiles of Ex, Ey, and Ez polarization used for training PINNs to retrieve the 3D permittivity profile. (d) Retrieved 3D permittivity profile for a sphere with radius r = 2 µm and ɛ = 3. The minimum size of the visualization grid is 0.5 µm and that of the computational grid is 200 × 200 × 333 nm3 along the x-, y-, and z-axes.

FIG. 10.

(a)–(c) Planar cross sections of the 3D electric field distribution (real part) profiles of Ex, Ey, and Ez polarization used for training PINNs to retrieve the 3D permittivity profile. (d) Retrieved 3D permittivity profile for a sphere with radius r = 2 µm and ɛ = 3. The minimum size of the visualization grid is 0.5 µm and that of the computational grid is 200 × 200 × 333 nm3 along the x-, y-, and z-axes.

Close modal

In conclusion, we have introduced a general DL framework for solving PDEs using PINNs to inversely retrieve unknown 2D and 3D electric and magnetic material parameters and shape information from synthetic field data. Our results are particularly interesting for inverse microscopy given the current availability of experimental near-field techniques that can measure the optical phase in the near zone with nanoscale resolution. By considering different complex PDE models and field data obtained from FEM simulations, we used PINNs to demonstrate successful retrieval of the complex ɛr(x, y) and μr(x, y) profiles simultaneously and with very good accuracy. We emphasize that this is achieved within the physics-informed method with significantly reduced data collection and training requirements compared to traditional machine learning approaches that typically employ massive datasets. We presented PINN-based parameter retrieval models that work under both extended and localized excitations that are typically used in SNOM applications. We then proposed and demonstrated an adaptive-PINN algorithm for improving the accuracy of the parameter retrieval for high-index materials. Finally, we showed a successful application of PINNs to the retrieval of the complex permittivity of a 3D scattering object with an unknown shape. The developed approach can be naturally scaled to any wavelength of interest and applied in arbitrary geometries, providing novel opportunities for non-invasive remote sensing techniques based on measured field data. The proposed computational framework can naturally enhance existing imaging techniques for the detection of magnetic nanoparticles used in cancer therapy and drug delivery38–40 and can be utilized for inspecting and characterizing complex optical devices based on acquired images.49 Although in this paper we were concerned with implementations of PINNs based on complex field data, phase retrieval techniques that recover the phase information from intensity measurements can be used to solve more general intensity-based (phase-less) retrieval problems with near-field imaging techniques.50–52 

The complex electric field and magnetic field data are obtained by solving the forward scattering problem using the finite element method.53 For the 2D examples, we used a minimum element size of 0.6 nm and a perfectly matching layer (PML) boundary with a thickness of 3 μm surrounding the square domain Ω of side length 10 µm. The total degrees of freedom of resulting FEM models are around 200 000. We implemented a scattered field formulation and set the background electric field as the plane wave propagating from left to right in the domain Ω. The complex field data are sampled on a 200 × 200 grid point in Ω.31 

The same minimum element size and PML boundary thickness are used for solving the 3D forward scattering problem. The degrees of freedom of the 3D FEM model are 700000. The incident plane wave propagates along the x-axis with the electric field polarized along the z-axis. We sampled the complex electric fields on a 3D grid with point numbers 50 × 50 × 30 along x-, y-, and z-axes, respectively. This provides a minimum size of the computational grid equal to 200 × 200 × 333 nm3 along x-, y-, and z-axes, respectively.

In all simulations except for the 3D retrieval case, a FCNN with four hidden layers and 64 neurons in each hidden layer is trained. We used a FCNN with three hidden layers with 20 neurons in each hidden layer for the 3D parameter retrieval. For all the PINNs discussed, we set the learning rate as 10−3. We fixed wi = 100 in the training process for a better convergence to the input field data for the standard PINNs. The adaptive PINNs used wf = 1, wb = 1, and wi = 100 as the initial loss weights. Notice that we use the same hyperparameter values when addressing different problems, which demonstrates the robustness of our methodology for solving parameter retrieval for microscopy. We choose the hyperbolic tangent function as the activation function. The Glorot uniform method is used for the ANN weights and bias initialization. The Adam optimizer is used for training the ANN.

The training process is implemented on a desktop with an Intel i7-8700K central processing unit (CPU) at 3.70 GHz and 32 Gb of RAM using a Nvidia GeForce GTX 1080Ti graphics processing unit (GPU). A typical training process for the FCNN takes around 10 h.

See the supplementary material for additional information.

L.D.N. acknowledges the support from the Army Research Laboratory (ARL; Cooperative Agreement No. W911NF-12-2-0023) and the National Science Foundation (Grant No. ECCS-2015700). The authors would like to thank Dr. Lu Lu and Professor George Em Karniadakis from Brown University for introducing us to the general methodology of PINNs and for insightful discussions.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
D.
Courjon
,
Near-Field Microscopy and Near-Field Optics
(
World Scientific Publishing Company
,
2003
).
2.
A. E.
Klein
,
N.
Janunts
,
S.
Schmidt
,
S.
Bin Hasan
,
C.
Etrich
,
S.
Fasold
,
T.
Kaiser
,
C.
Rockstuhl
, and
T.
Pertsch
, “
Dual-SNOM investigations of multimode interference in plasmonic strip waveguides
,”
Nanoscale
9
,
6695
6702
(
2017
).
3.
S. G.
Stanciu
,
D. E.
Tranca
,
L.
Pastorino
,
S.
Boi
,
Y. M.
Song
,
Y. J.
Yoo
,
S.
Ishii
,
R.
Hristu
,
F.
Yang
,
G.
Bussetti
 et al., “
Characterization of nanomaterials by locally determining their complex permittivity with scattering-type scanning near-field optical microscopy
,”
ACS Appl. Nano Mater.
3
,
1250
1262
(
2020
).
4.
W.
Zhang
and
Y.
Chen
, “
Visibility of subsurface nanostructures in scattering-type scanning near-field optical microscopy imaging
,”
Opt. Express
28
,
6696
6707
(
2020
).
5.
L.
Novotny
and
B.
Hecht
,
Principles of Nano-Optics
(
Cambridge University
,
2012
).
6.
E. A.
Ash
and
G.
Nicholls
, “
Super-resolution aperture scanning microscope
,”
Nature
237
,
510
512
(
1972
).
7.
E.
Betzig
and
J. K.
Trautman
, “
Near-field optics: Microscopy, spectroscopy, and surface modification beyond the diffraction limit
,”
Science
257
,
189
195
(
1992
).
8.
R.
Dallapiccola
,
C.
Dubois
,
A.
Gopinath
,
F.
Stellacci
, and
L.
Dal Negro
, “
Near-field excitation and near-field detection of propagating surface plasmon polaritons on Au waveguide structures
,”
Appl. Phys. Lett.
94
,
243118
(
2009
).
9.
A. A.
Govyadinov
,
S.
Mastel
,
F.
Golmar
,
A.
Chuvilin
,
P. S.
Carney
, and
R.
Hillenbrand
, “
Recovery of permittivity and depth from near-field data as a step toward infrared nanotomography
,”
ACS Nano
8
,
6911
6921
(
2014
).
10.
A. E.
Klein
,
N.
Janunts
,
M.
Steinert
,
A.
Tünnermann
, and
T.
Pertsch
, “
Polarization-resolved near-field mapping of plasmonic aperture emission by a dual-SNOM system
,”
Nano Lett.
14
,
5010
5015
(
2014
).
11.
B.
Chen
and
J. J.
Stamnes
, “
Validity of diffraction tomography based on the first Born and the first Rytov approximations
,”
Appl. Opt.
37
,
2996
3006
(
1998
).
12.
U. S.
Kamilov
,
I. N.
Papadopoulos
,
M. H.
Shoreh
,
A.
Goy
,
C.
Vonesch
,
M.
Unser
, and
D.
Psaltis
, “
Learning approach to optical tomography
,”
Optica
2
,
517
522
(
2015
).
13.
S.
Molesky
,
Z.
Lin
,
A. Y.
Piggott
,
W.
Jin
,
J.
Vucković
, and
A. W.
Rodriguez
, “
Inverse design in nanophotonics
,”
Nat. Photonics
12
,
659
670
(
2018
).
14.
H.-Y.
Liu
,
D.
Liu
,
H.
Mansour
,
P. T.
Boufounos
,
L.
Waller
, and
U. S.
Kamilov
, “
SEAGLE: Sparsity-driven image reconstruction under multiple scattering
,”
IEEE Trans. Comput. Imaging
4
,
73
86
(
2017
).
15.
U. S.
Kamilov
,
H.
Mansour
, and
B.
Wohlberg
, “
A plug-and-play priors approach for solving nonlinear imaging inverse problems
,”
IEEE Signal Process. Lett.
24
,
1872
1876
(
2017
).
16.
T.-A.
Pham
,
E.
Soubies
,
A.
Goy
,
J.
Lim
,
F.
Soulez
,
D.
Psaltis
, and
M.
Unser
, “
Versatile reconstruction framework for diffraction tomography with intensity measurements and multiple scattering
,”
Opt. Express
26
,
2749
2763
(
2018
).
17.
D.
Colton
and
R.
Kress
, “
Looking back on inverse scattering theory
,”
SIAM Rev.
60
,
779
807
(
2018
).
18.
Z.
Wei
and
X.
Chen
, “
Deep-learning schemes for full-wave nonlinear inverse scattering problems
,”
IEEE Trans. Geosci. Remote Sens.
57
,
1849
1860
(
2019
).
19.
Y.
Sun
,
Z.
Xia
, and
U. S.
Kamilov
, “
Efficient and accurate inversion of multiple scattering with deep learning
,”
Opt. Express
26
,
14678
14688
(
2018
).
20.
Y.
Sanghvi
,
Y.
Kalepu
, and
U. K.
Khankhoje
, “
Embedding deep learning in inverse scattering problems
,”
IEEE Trans. Comput. Imaging
6
,
46
(
2019
).
21.
J.
Lim
,
A. B.
Ayoub
, and
D.
Psaltis
, “
Three-dimensional tomography of red blood cells using deep learning
,”
Adv. Photonics
2
,
1
9
(
2020
).
22.
S.
Mastel
,
A. A.
Govyadinov
,
C.
Maissen
,
A.
Chuvilin
,
A.
Berger
, and
R.
Hillenbrand
, “
Understanding the image contrast of material boundaries in IR nanoscopy reaching 5 nm spatial resolution
,”
ACS Photonics
5
,
3372
3378
(
2018
).
23.
W.
Ma
,
Z.
Liu
,
Z. A.
Kudyshev
,
A.
Boltasseva
,
W.
Cai
, and
Y.
Liu
, “
Deep learning for the design of photonic structures
,”
Nat. Photonics
15
,
77
90
(
2021
).
24.
Z.
Liu
,
D.
Zhu
,
L.
Raju
, and
W.
Cai
, “
Tackling photonic inverse design with machine learning
,”
Adv. Sci.
8
,
2002923
(
2021
).
25.
I.
Tanriover
,
W.
Hadibrata
, and
K.
Aydin
, “
Physics-based approach for a neural networks enabled design of all-dielectric metasurfaces
,”
ACS Photonics
7
,
1957
1964
(
2020
).
26.
I.
Tanriover
,
W.
Hadibrata
,
J.
Scheuer
, and
K.
Aydin
, “
Neural networks enabled forward and inverse design of reconfigurable metasurfaces
,”
Opt. Express
29
,
27219
27227
(
2021
).
27.
T.
Repän
,
R.
Venkitakrishnan
, and
C.
Rockstuhl
, “
Artificial neural networks used to retrieve effective properties of metamaterials
,”
Opt. Express
29
,
36072
36085
(
2021
).
28.
C. C.
Nadell
,
B.
Huang
,
J. M.
Malof
, and
W. J.
Padilla
, “
Deep learning for accelerated all-dielectric metasurface design
,”
Opt. Express
27
,
27523
27535
(
2019
).
29.
R.
Tempke
,
L.
Thomas
,
C.
Wildfire
,
D.
Shekhawat
, and
T.
Musho
, “
Machine learning approach to transform scattering parameters to complex permittivities
,”
J. Microwave Power Electromagn. Energy
55
(
4
),
287
302
(
2021
).
30.
M.
Raissi
,
P.
Perdikaris
, and
G. E.
Karniadakis
, “
Physics-informed neural networks: A deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations
,”
J. Comput. Phys.
378
,
686
707
(
2019
).
31.
L.
Lu
,
X.
Meng
,
Z.
Mao
, and
G. E.
Karniadakis
, “
DeepXDE: A deep learning library for solving differential equations
,”
SIAM Rev.
63
,
208
228
(
2021
).
32.
G.
Pang
,
L.
Lu
, and
G. E.
Karniadakis
, “
fPINNs: Fractional physics-informed neural networks
,”
SIAM J. Sci. Comput.
41
,
A2603
A2626
(
2019
).
33.
Y.
Chen
,
L.
Lu
,
G. E.
Karniadakis
, and
L.
Dal Negro
, “
Physics-informed neural networks for inverse problems in nano-optics and metamaterials
,”
Opt. Express
28
,
11618
11633
(
2020
).
34.
R. W.
Boyd
,
Nonlinear Optics
(
Academic Press
,
2020
).
35.
M.
Abadi
,
P.
Barham
,
J.
Chen
,
Z.
Chen
,
A.
Davis
,
J.
Dean
,
S.
Ghemawat
,
G.
Irving
,
M.
Isard
 et al., “
Tensorflow: A system for large-scale machine learning
,” in
OSDI’16: 12th USENIX Symposium on Operating Systems Design and Implementation
(
OSDI
,
2016
), pp.
265
283
.
36.
G. E.
Karniadakis
,
I. G.
Kevrekidis
,
L.
Lu
,
P.
Perdikaris
,
S.
Wang
, and
L.
Yang
, “
Physics-informed machine learning
,”
Nat. Rev. Phys.
3
,
422
440
(
2021
).
37.
J. D.
Jackson
,
Classical Electrodynamics
, 3rd ed. (
Wiley
,
1998
).
38.
X.
Li
,
J.
Wei
,
K. E.
Aifantis
,
Y.
Fan
,
Q.
Feng
,
F.-Z.
Cui
, and
F.
Watari
, “
Current investigations into magnetic nanoparticles for biomedical applications
,”
J. Biomed. Mater. Res., Part A
104
,
1285
1296
(
2016
).
39.
L.
Mohammed
,
H. G.
Gomaa
,
D.
Ragab
, and
J.
Zhu
, “
Magnetic nanoparticles for environmental and biomedical applications: A review
,”
Particuology
30
,
1
14
(
2017
).
40.
V. F.
Cardoso
,
A.
Francesko
,
C.
Ribeiro
,
M.
Bañobre-López
,
P.
Martins
, and
S.
Lanceros-Mendez
, “
Advances in magnetic nanoparticles for biomedical applications
,”
Adv. Healthcare Mater.
7
,
1700845
(
2018
).
41.
I. S.
Maksymov
, “
Magneto-plasmonic nanoantennas: Basics and applications
,”
Rev. Phys.
1
,
36
51
(
2016
).
42.
K.
Shukla
,
P. C.
Di Leoni
,
J.
Blackshire
,
D.
Sparkman
, and
G. E.
Karniadakis
, “
Physics-informed neural network for ultrasound nondestructive quantification of surface breaking cracks
,”
J. Nondestr. Eval.
39
,
61
(
2020
).
43.
C.
Bajaj
,
L.
McLennan
,
T.
Andeen
, and
A.
Roy
, “
Robust learning of physics informed neural networks
,” arXiv:2110.13330 (
2021
).
44.
G. Y.
Panasyuk
,
V. A.
Markel
,
P.
Scott Carney
, and
J. C.
Schotland
, “
Nonlinear inverse scattering and three-dimensional near-field optical imaging
,”
Appl. Phys. Lett.
89
,
221116
(
2006
).
45.
G.
Bao
and
P.
Li
, “
Inverse medium scattering problems in near-field optics
,”
J. Comput. Math.
25
,
252
265
(
2007
).
46.
L. D.
McClenny
and
U. M.
Braga-Neto
, “
Self-adaptive physics-informed neural networks using a soft attention mechanism
,” arXiv:2009.04544 (
2020
).
47.
C. L.
Wight
and
J.
Zhao
, “
Solving Allen-Cahn and Cahn-Hilliard equations using the adaptive physics informed neural networks
,”
Commun. Comput. Phys.
29
(
3
),
930
954
(
2021
).
48.
S.
Wang
,
Y.
Teng
, and
P.
Perdikaris
, “
Understanding and mitigating gradient pathologies in physics-informed neural networks
,” arXiv:2001.04536 (
2020
).
49.
M.
Kadic
,
G. W.
Milton
,
M.
van Hecke
, and
M.
Wegener
, “
3D metamaterials
,”
Nat. Rev. Phys.
1
,
198
210
(
2019
).
50.
G.
Gbur
and
E.
Wolf
, “
Diffraction tomography without phase information
,”
Opt. Lett.
27
,
1890
1892
(
2002
).
51.
L.
Tian
and
L.
Waller
, “
3D intensity and phase imaging from light field measurements in an LED array microscope
,”
Optica
2
,
104
111
(
2015
).
52.
H.
Ammari
,
Y. T.
Chow
, and
J.
Zou
, “
Phased and phaseless domain reconstructions in the inverse scattering problem via scattering coefficients
,”
SIAM J. Appl. Math.
76
,
1000
1030
(
2016
).
53.
COMSOL AB
, Stockholm, Sweden, COMSOL Multiphysics® v.5.4, www.comsol.com.

Supplementary Material