Abstract
The ability to perform mathematical computations using metastructures is an emergent paradigm that carries the potential of wave-based analog computing to the realm of near-speed-of-light, low-loss, compact devices. We theoretically introduce and experimentally verify the concept of a reconfigurable metastructure that performs analog complex mathematical computations using electromagnetic waves. Reconfigurable, RF-based components endow our device with the ability to perform stationary and non-stationary iterative algorithms. After demonstrating matrix inversion (stationary problem), we use the machine to tackle two major non-stationary problems: root finding with Newtonâs method and inverse design (constrained optimization) via the Lagrange multiplier method. The platform enables possible avenues for wave-based, analog computations for general linear algebraic problems and beyond in compact, ultrafast, and parallelized ways.
Similar content being viewed by others
Introduction
Calculators of various kinds have emerged by forging numerical algorithms with the corresponding technological platforms. While the algorithms describe the mathematical paths on how solutions to problems can be found, the platforms are responsible for the transliteration of this abstract path into measurable quantities. The algorithms, the platforms, and their fusion define such systemsâ features and limitations. Following the ever-growing demand for ultrafast, compact, low/near-zero-power, and integrable cyber-physical devices for mathematical computations, it is organic that significant research efforts focus on making these numerical systems as optimal and efficient as possible.
This quest led to the exploration and development of unconventional analog computing systems that exploit electromagnetic waves to deliver low-power, ultrafast, compact, parallelized computations1,2,3,4,5,6,7. The two main categories in this domain involve systems that utilize free-space (scattering) elements (e.g., lenses6), and waveguides (e.g. photonic systems8,9 and phased arrays10). Sufficient free-space propagation can act as dense matrix multiplication11. Realized with traditional optics, this results in bulky devices6,12, while metasurfaces can be more compact13,14,15. However, in both, there can be major bottlenecks regarding photonic and electronic integration. Waveguiding systems such as photonic integrated circuits (IC) offer more mature solutions for integrable and reconfigurable devices at the expense of the challenge of rendering all of the needed functionality in a 2D plane16. In all cases, their main challenge is reconfigurability, since its implementation requires some form of a-priori mathematical calculations. For instance, metasurfaces/complex media17,18 require optimization, and photonic meshes require operator decomposition19,20,21,22.
In terms of their mathematical abilities, the above examples demonstrate wave-based analog computing with functionalities such as integration/differentiation in space23,24,25,26 and time27, matrix-vector multiplication28, emulating equations through physical phenomena29,30, or acting as platforms for neural network functionalities3,6,9,31. The intersection with the metamaterial paradigm delivered a series of remarkable analog computing devices with matrix multiplication1,7,23 and ultimately equation-solving (matrix inversion) capabilities2. In most cases, the matrix computations (especially the matrix inversion2) were performed through stationary algorithms32, such as the Jacobi method, where the matrix (operator/kernel) does not change with the iteration count.
The fundamental and far-reaching question we address here is whether a wave-based analog metastructure can be reconfigurable simply and intuitively, without needing a-priori calculations. Most importantly, the resolution to this question endows one with the ability to implement stationary and non-stationary algorithms. We propose a device based on an RF waveguide architecture with reconfigurable components. Regarding stationary problems, we use this device to perform a matrix inversion of a random (statistically uncorrelated) large set of matrices. As for non-stationary problems, we demonstrate both root finding using Newtonâs method and Inverse Design. All three examples are possible only due to the reconfigurability of the device and hint at the possibility of deeper explorations into the realm of advanced numerical algebra methods.
Results
A conceptual representation of the main idea is pictorially summarized in Fig. 1A. The main property of our proof-of-concept system distinctively different from all previous metastructure approaches (e.g.,2) is its reconfigurability; the metastructure has the ability to rapidly take on different matrices (operators or kernels) K. To facilitate this, we employ a wave-based direct complex matrix (DCM) architecture, which offers an intuitive and simple implementation of any desired matrix33. Using waves instead of currents and voltages, it is analogous to the crossbar architecture used in electronic analog computing systems34,35,36,37, and it can be seen as a generalized phased array feed10. In this device, a collection of nâÃân tunable phase shifting and amplifying elements (which can also act as attenuating elements) connect an input vector of n complex amplitudes on an array of transmission lines to a similar output vector through combiners. This architecture can be seen schematically in Fig. 1B and its corresponding experimental implementation in Fig. 1C.
The key component of the metadevice is the multiplier module (Fig. 1D) so named because given an input signal characterized by its complex amplitude at 45âMHz, Vin, it will render a similar output Voutâ=âzVin, where z is a complex multiplication factor. This module consists of two basic components: (a) a voltage-controlled phase shifter with over 360 degrees of potential phase rotation, and (b) a voltage-controlled amplifier with â47âdB of dynamic range (â30âdB to +17âdB). Each multiplier contains an embedded microcontroller unit (MCU), which allows each device to be controlled externally through a suitable communication network and a computer (see supplementary material). While the design frequency is 45âMHz, the module could be implemented at RF (GHz) and photonic (THz) platforms, following the same principle of operation. The experimental DCM implementation consists of 25 multipliers to yield 5âÃâ5 complex matrices. The ingress and egress stages33 are implemented with five 1-to-5 power splitters (ingress stage) and five 5-to-1 signal combiners (egress stage). The multipliers are clustered into five groups, one for each matrix row. In Fig. 1B we depict the planar schematic of the DCM suitable for photonic implementation. However, for the RF implementation, we stacked and routed the components vertically (Fig. 1C), making the device compact for our particular wavelength and platform choice. Different stacking or integrated circuitry approaches can potentially be used to further reduce its overall footprint.
The metadevice can be operated in one of two configurations with dramatically different results. When the DCM is set in an open-loop configuration (Fig. 1E inset), it can be used to rapidly calculate parallelized matrix-vector multiplication. However, a closed-loop configuration can be created by connecting the outputs and the inputs with a feedback loop using properly designed couplers (Fig. 1F inset). When the DCM is in a closed-loop configuration, the metadevice can rapidly calculate parallelized matrix inversion (equation solving). This is a unique feature of metadevices/metastructures2,33 that incorporate feedback loops.
Inverting stationary matrices
First, we investigate the stationary analysis capabilities of our metadevice. For the assessment of the open- and closed-loop operation, we performed a series of randomized trials, one instance of which is presented in Fig. 1E, F. For each trial, a random passive matrix \(A\in {{\mathbb{C}}}^{5\times 5}\) was chosen and applied to the metadevice in both configurations. Each measurement was performed by exciting each input port in turn with all other inputs appropriately terminated and then observing the complex amplitudes on each of the output ports. For the open-loop configuration, this corresponds to performing five matrix-vector multiplications, or as A â I where each column of I is progressively applied (one column at a time) as separate vectors. The closed-loop configuration was measured similarly, but in this case, we are probing the steady state of the metadevice, which corresponds to (IâââK)â1 â Iâ=âAâ1 â I. While this measurement technique fully characterizes both configurations, in practice dense complex vectors will be input and read to achieve parallelized results.
The estimated relative error â¥AexactâââAmeasâ¥2/â¥Aexactâ¥2 for both cases (Fig. 1E, F) revealed an error of about 0.001 and 0.005, respectively. Despite the component imperfections, misalignments, measurement noise, and other stochastic errors, the measured results are in excellent agreement with the theoretical values. The calibration procedure of the metadevice and the statistical analysis of the full trial set (100 values) are presented in the Supplementary Material.
A single multiplier module has a rise time of approximately 80âns to achieve its desired complex value. This value is approximately 4T assuming one-period duration of Tâ=â1/45âMHz â 22.2âns. In the open-loop configuration, the total response requires approximately 5T, including signal delays in connections and splitters. The duration of the closed-loop case is affected by the platform and the condition number of the inverted matrix33, but in principle is in the same order of magnitude. Possible photonic implementations may further reduce this time to the picosecond range3 and below38.
Root finding with Newtonâs method
We now apply the implemented metadevice to two characteristic non-stationary problems that highlight its mathematical abilities: (i) root finding of a system of five equations with five unknowns using Newtonâs iterative technique and (ii) implementing an inverse design problem using the Lagrangian multiplier formalism for constrained optimization. Both cases require that the kernel be reprogrammed in each iteration step. Note that our approach is not restricted to these two problems; instead, we choose these to highlight the potential of the introduced metadevice.
For the first case, we construct a simple nonlinear toy problem, and we apply Newtonâs algorithm39 (Fig. 2A) to find one possible root. The vector problem statement reads
where \({{\bf{f}}}\in {{\mathbb{C}}}^{5\times 1}\), \({{\bf{z}}}={\left[{z}_{1},{z}_{2},{z}_{3},{z}_{4},{z}_{5}\right]}^{T}\in {{\mathbb{C}}}^{5\times 1}\), and 0 is the zero vector. We construct the vector function to have the following polynomial form
where \({{\bf{r}}}={\left[{r}_{1},{r}_{2},{r}_{3},{r}_{4},{r}_{5}\right]}^{T}\) are the vertices of a regular pentagon with 1/4 radius (see Fig. 2B) and the other factors represent additional extraneous roots far from the starting point.
For the evaluation of Newtonâs method, we need to calculate the Jacobian matrix, i.e., \({J}_{ij}=\frac{\partial {f}_{i}}{\partial {z}_{j}}\) or
therefore the root can be estimated by the following iterative process
where αâ=â0.2 is a relaxation constant33.
Table 1 shows the required algorithm steps to implement the iterative scheme described by Eq. (8). Note that the Jacobian changes value in each iteration and that its inverse is required to be calculated anew. The results are then used to update the z. The method converges successfully after a few iterations.
Matrix inversion represents a computationally intensive task, expedited by our metadevice. In Newtonâs method, matrix inversion stands as the primary complexity, with an \({{\mathcal{O}}}({n}^{3})\) complexity assumption for an nâÃân matrix40. On the other hand, evaluating the Jacobian typically demands \({{\mathcal{O}}}(n)\) operations, facilitated by automatic differentiation techniques29. The efficient computation of the Jacobian at each step can be provided to the system either from a dedicated subsystem responsible for its calculation and feed into the main system.
A numerical version (using MATLAB) is compared with the experimental results illustrated in Fig. 2B. We observe that for both MATLAB and the experiment, the estimation vector converges close to the exact roots. Furthermore, the estimated vector reaches a stationary point as the iteration count increases. After 15 iterations, the relative error is â¥zââârâ¥2/â¥râ¥2âââ0.0023. This is similar to the accuracy achieved for stationary trials, thus representing the accuracy floor of our system. A similar picture is also visible by comparing three specific iterations, as illustrated in Fig. 2C, where a comparison of the full Jacobian is presented.
The experimental results do not precisely follow the paths indicated by the numerical implementation realized using MATLAB. This can be explained by adding random noise to the Jacobian on each iteration step. The added noise \(N\in {{\mathbb{C}}}^{5\times 5}\) is a random complex matrix that follows a normal distribution inside a disk with radius \({r}_{N}=0.01{\lambda }_{\max }\), where \({\lambda }_{\max }\) is the maximum eigenvalue of the Jacobian. The noise creates many possible paths, all of which successfully converge, and we observe that our measured results comfortably lie within these families of curves. Note that some solution branches are more susceptible to this noise than others (e.g., r1 (blue) and r3 (red) curves in Fig. 2B) and this is due to the details of the toy problem solved.
Generally, the deviceâs numerical accuracy has a threshold that depends both on the implementation and on the measuring apparatus (vector network analyzer (VNA)). In our case, the closed-loop operation was on the order of a few microvolts, approaching the sensitivity floor of the VNA utilized (refer to Supplementary Material for details). Additionally, the implementation would influence the scalability of systems handling matrices with tens or hundreds of inputs. Thoughtful RF/Photonic IC designs might pave the way for incorporating larger kernels, albeit requiring more sensitive VNAs. It is crucial to emphasize phase and amplitude stability, as highlighted in ref. 33.
When higher-precision computations are required, this device can be a part of a mixed-precision computing system. In these systems, part of the calculations are done in a fast, low-precision estimation stage and then fed and further refined at a higher-precision stage, similar to the in-memory mixed-precision approaches in electronic platforms41.
Inverse design of a metastructure
For the second example, we chose the case of an inverse design problem (Fig. 3A). We assume that our design consists of a collection of mâ=â5 two-dimensional (2D) scatterers with circular cross-section at fixed known locations râ=â[r1, â¦, r5], each with an unknown bounded permittivity \({{\mathbf{\varepsilon }}}=[{\varepsilon }_{1},\ldots,{\varepsilon }_{5}]\in {{\mathbb{C}}}^{5\times 1}\). The goal is to achieve a specific user-defined scattered field measured at a series of nâ=â4 detection (objective) points, oâ=â[o1, â¦, o4]. Note that in our case we assume a collection of cylindrical circular scatterers (2D) excited with a monochromatic incident field of λw wavelength. The x-propagating incident field (kx) is polarized in the z-direction (TE waveâ-âEz) with the ejÏt convention.
The scatterers are coupled, making this a nonlinear problem modeled using the LippmannâSchwinger42 scheme, solved with a standard discrete dipole approximation (DDA) methodology43. Each scatterer will respond to the local (self-excluded) electric field, which consists of the known incident electric field, \({e}_{{{\rm{inc}}}}\in \,{{\mathbb{C}}}^{5\times 1}\), and the scattered field from all other scatterers, \({e}_{{{\rm{sca}}}}\in \,{{\mathbb{C}}}^{5\times 1}\). The scatterers exhibit a complex polarization vector \(p=A({e}_{{{\rm{inc}}}}+{e}_{{{\rm{sca}}}})\in \,{{\mathbb{C}}}^{5\times 1}\) where A is the normalized polarizability diagonal matrix, i.e., Aâ=âdiag(εâââεbackground). The field interaction between the scatterers is expressed via the Green matrix \(G\in \,{{\mathbb{C}}}^{5\times 5}\) (hollow symmetric matrix) such that escaâ=âGp. We may express the polarization vector pâ=âA(eincâ+âGp), which indicates the mutual dependence of p. Therefore, the polarization vector can be calculated as \(p={({A}^{-1}-G)}^{-1}{e}_{{{\rm{inc}}}}\). Finally, we use the four objective points o to measure the scattered field vector \({e}_{{{\rm{meas}}}}={G}_{{{\rm{pr}}}}p\in \,{{\mathbb{C}}}^{4\times 1}\) where \({G}_{{{\rm{pr}}}}\in {{\mathbb{C}}}^{4\times 5}\) is the propagator Green function. The measured field is then compared to a (user-defined) objective \({e}_{{{\rm{obj}}}}\in \,{{\mathbb{C}}}^{4\times 1}\).
A typical constrained minimization problem (primal) can be written as39
where f(x, y) are the objectives and g(x, y) are the constraints. For such problems, the Lagrangian (dual) problem is expressed as
Note that x and y may be subject to further requirements such as domains and bounds.
For our particular example, we have that xâ=âp, yâ=âε, and f(p, ε)â=â1/2â¥Gprpâââeobjâ¥2 and g(p, ε)â=â1/2â¥(A(ε)â1âââG)pâââeincâ¥2. In our formulation, the objective is the scattered field at the observation points. The constraints comprise the self-consistency of the polarization vector (physics). Also, the permittivity vector is subject to specific bounds, i.e., \(\varepsilon \in {\mathbb{R}}\) and ε â [1, 5]. Note that g(p, ε) is nonlinear with respect to p and ε and therefore requires a non-stationary approach.
Following an initialization, our numerical evaluation of the above is implemented by a non-stationary algorithm that requires repeated application of the following three steps. First, we minimize with respect to ε by examining \({\nabla }_{\varepsilon }{{\mathcal{L}}}(p,\varepsilon,\lambda )=0\). At this step, we project the resulting permittivity vector to the desired domain and bounds. Second, we minimize with respect to p by examining \({\nabla }_{p}{{\mathcal{L}}}(p,\varepsilon,\lambda )=0\). At this stage, the required stationary matrix inversion is performed with our metadevice. Finally, we maximize for λ by using \({\nabla }_{\lambda }{{\mathcal{L}}}(p,\varepsilon,\lambda )=0\). These steps are repeated until convergence is achieved, i.e.,
(for more information, see SM).
As a numerical test case, the scatterers are assumed to be lossless with permittivity of \(\varepsilon=\left[{\varepsilon }_{1},{\varepsilon }_{2},{\varepsilon }_{3},{\varepsilon }_{4},{\varepsilon }_{5}\right]=[3.5,\,1.5,\,1.5,\,3.5,\,1.5]\). The objective scattered field at the detection points o, as depicted in (Fig. 3A), is eobjâ=â[â0.0086âââ0.0078j, 0.0089âââ0.0132j,â0.0066âââ0.0120j, 0.0043âââ0.0004j]. The values were extracted from the DDA method and verified with a full-wave COMSOL simulation. Note that Fig. 3A depicts the complex (hue/saturation) of the electric scattered field (Ez), i.e., the difference between the total field and the incident excitation.
Figure 3B depicts a set of four cases for the same algorithm. In the first case (black line), the idealized (noiseless, no filtering) computer evaluation of the algorithm is given - we observe that after only 20 iterations the error drops below 10â3. The experimental results are presented in Fig. 3B as red dots. The measured results exhibit an optimal point (minimum error) after 87 iterations, with an error of 0.00172. As an analog device, there is an additional systematic/stochastic/experimental noise to the system which affects the fidelity of the matrix inversion. We apply a simple averaging filtering scheme on the polarization estimation, i.e., pnewâ=â(1âââαF)pâ+âαFpprevious, with αFâ=â0.25, as a way to partially mitigate this noise. The filter affects the convergence speed by increasing the iteration count but also significantly improves the accuracy/fidelity of the matrix inversion, hence the metadeviceâs performance. This feature is illustrated in Fig. 3B, where the retrieved experimental results are compared to the idealized computer evaluation with the applied filter (blue line). We also performed a series of 100 randomized cases of the idealized filtered computer evaluation with added noise to the estimated/measured polarization vector (faint blue lines in Fig. 3B). The noise profile is similar to the one used in the first example (Newtonâs method). The measured results are well contained within these error bounds. Note that iteration count is not equivalent to time. For a traditional computer evaluation, each iteration (with its required matrix inversion) could ultimately be slower than the convergence time of an optimized hardware implementation of the metadevice.
Due to systematic/measurement noise, the error begins to grow after the experimental accuracy floor is obtained - an indication that a termination criterion could be applied at this point. This result also agrees with the maximum accuracy we obtained in the previous non-stationary example. More sophisticated error-correcting and filtering schemes can push the accuracy below this threshold. For instance, αF could be adaptively tuned during the non-stationary evaluation to realize a mixed-precision computing system.
At the minimum error point (iteration 87), the extracted permittivity estimation is illustrated in (Fig. 3C). Notice that the values are very close to the numerical test case objectives and permittivities. Finally, Fig. 3D illustrates the path of the scattering vector, emeas, for these 87 iterations. Similar to the above example, the faint paths represent the added noise effects on the numerical evaluation.
Discussion
For both presented non-stationary examples, it is evident that our metadevice can act either as an ultrafast analog computing machine and mathematics calculator with waves, or in a broader sense as an electromagnetic emulator for inverse design44. It can be used for a plethora of realistic problems where the linear response of a system (i.e. matrix-vector multiplication) or the solution of a system of equations (stationary problems, matrix inversion) is required. Active control systems45, real-time signal processing46, and radar systems47 are a few indicative systems that require real-time analog processing. Moreover, the intuitive reconfigurability of this metadevice also enables the performance of constrained optimization tasks, like the ones required in non-stationary problems such as inverse design, where the desired response of complex media requires intensive optimization48. In short, this metastructure can design metastructures. Finally, an adaptation of the above proof-of-concept metadevice in RF-IC, photonic, or hybrid platforms can make it an excellent candidate for on-the-fly or computation-through-propagation ultrafast, parallelized calculations. Scaling such metastructures to larger matrices is an open challenge.
Data availability
All data and materials needed to evaluate the results and conclusions in the paper are present in the main text and the supplementary materials. The raw data and other relevant materials are available at https://doi.org/10.6084/m9.figshare.26936227.v1.
References
Silva, A. et al. Performing mathematical operations with metamaterials. Science 343, 160â163 (2014).
Mohammadi Estakhri, N., Edwards, B. & Engheta, N. Inverse-designed metastructures that solve equations. Science 363, 1333â1338 (2019).
Ashtiani, F., Geers, A. J. & Aflatouni, F. An on-chip photonic deep neural network for image classification. Nature 606, 501â506 (2022).
Caulfield, H. J. & Dolev, S. Why future supercomputing requires optics. Nat. Photonics 4, 261â263 (2010).
Solli, D. R. & Jalali, B. Analog optical computing. Nat. Photonics 9, 704â706 (2015).
Wetzstein, G. et al. Inference in artificial intelligence with deep optics and photonics. Nature 588, 39â47 (2020).
Zangeneh-Nejad, F., Sounas, D. L., Alù, A. & Fleury, R. Analogue computing with metamaterials. Nat. Rev. Mater. 6, 207â225 (2021).
Feldmann, J. et al. Parallel convolutional processing using an integrated photonic tensor core. Nature 589, 52â58 (2021).
TeÄin, U., YÃldÃrÃm, M., OÄuz, I., Moser, C. & Psaltis, D. Scalable optical learning operator. Nat. Comput. Sci. 1, 542â549 (2021).
Sun, J., Timurdogan, E., Yaacobi, A., Hosseini, E. S. & Watts, M. R. Large-scale nanophotonic phased array. Nature 493, 195â199 (2013).
Goodman, J. W. Introduction to Fourier optics 3rd edn (Roberts & Co., Englewood, CO, 2005).
Vellekoop, I. M. & Mosk, A. P. Focusing coherent light through opaque strongly scattering media. Opt. Lett. 32, 2309â2311 (2007).
Shaltout, A. M., Shalaev, V. M. & Brongersma, M. L. Spatiotemporal light control with active metasurfaces. Science 364, eaat3100 (2019).
Imani, M. F. et al. Review of metasurface antennas for computational microwave imaging. IEEE Trans. Antennas Propag. 68, 1860â1875 (2020).
Cheben, P., Halir, R., Schmid, J. H., Atwater, H. A. & Smith, D. R. Subwavelength integrated photonics. Nature 560, 565â572 (2018).
Nikkhah, V. et al. Inverse-designed low-index-contrast structures on a silicon photonics platform for vector-matrix multiplication. Nat. Photonics, 18, 501â508 (2024).
Matthès, M. W., del Hougne, P., de Rosny, J., Lerosey, G. & Popoff, S. M. Optical complex media as universal reconfigurable linear operators. Optica 6, 465 (2019).
Venkatesh, S., Lu, X., Saeidi, H. & Sengupta, K. A programmable terahertz metasurface with circuit-coupled meta-elements in silicon chips: creating low-cost, large-scale, reconfigurable terahertz metasurfaces. IEEE Antennas Propag. Mag. 64, 110â122 (2022).
Reck, M., Zeilinger, A., Bernstein, H. J. & Bertani, P. Experimental realization of any discrete unitary operator. Phys. Rev. Lett. 73, 58â61 (1994).
Miller, D. A. B. Self-configuring universal linear optical component [Invited]. Photonics Res. 1, 1â15 (2013).
Clements, W. R., Humphreys, P. C., Metcalf, B. J., Kolthammer, W. S. & Walmsley, I. A. Optimal design for universal multiport interferometers. Optica 3, 1460â1465 (2016).
Marpaung, D., Yao, J. & Capmany, J. Integrated microwave photonics. Nat. Photonics 13, 80â90 (2019).
Pors, A., Nielsen, M. G. & Bozhevolnyi, S. I. Analog computing using reflective plasmonic metasurfaces. Nano Lett. 15, 791â797 (2015).
Zhu, T. et al. Plasmonic computing of spatial differentiation. Nat. Commun. 8, 15391 (2017).
Cordaro, A. et al. High-index dielectric metasurfaces performing mathematical operations. Nano Lett. 19, 8418â8423 (2019).
Zhou, Y., Zheng, H., Kravchenko, I. I. & Valentine, J. Flat optics for image differentiation. Nat. Photonics 14, 316â323 (2020).
Zangeneh-Nejad, F. & Fleury, R. Topological analog signal processing. Nat. Commun. 10, 2058 (2019).
Macho-Ortiz, A., Pérez-López, D. & Capmany, J. Optical implementation of 2 à 2 universal unitary matrix transformations. Laser Photon. Rev. 15, 2000473 (2021).
Hughes, T. W., Williamson, I. A. D., Minkov, M. & Fan, S. Wave physics as an analog recurrent neural network. Sci. Adv. 5, eaay6946 (2019).
Vadlamani, S. K., Xiao, T. P. & Yablonovitch, E. Physics successfully implements Lagrange multiplier optimization. Proc. Natl Acad. Sci. USA 117, 26639â26650 (2020).
Lin, X. et al. All-optical machine learning using diffractive deep neural networks. Science 361, 1004â1008 (2018).
Barrett, R. et al. Templates for the Solution of Linear Systems: Building Blocks for Iterative Methods (Society for Industrial and Applied Mathematics, 1994).
Tzarouchis, D. C., Mencagli, M. J., Edwards, B. & Engheta, N. Mathematical operations and equation solving with reconfigurable metadevices. Light Sci. Appl. 11, 263 (2022).
Ielmini, D. & Wong, H.-S. P. In-memory computing with resistive switching devices. Nat. Electron. 1, 333â343 (2018).
Zidan, M. A. et al. A general memristor-based partial differential equation solver. Nat. Electron. 1, 411â420 (2018).
Sun, Z. et al. Solving matrix equations in one step with cross-point resistive arrays. Proc. Natl Acad. Sci. USA 116, 4123â4128 (2019).
Mourgias-Alexandris, G. et al. Noise-resilient and high-speed deep learning with coherent silicon photonics. Nat. Commun. 13, 5572 (2022).
Guo, Q. et al. Femtojoule femtosecond all-optical switching in lithium niobate nanophotonics. Nat. Photonics 16, 625â631 (2022).
Bertsekas, D. P. Nonlinear Programming (Athena Scientific, Belmont, MA, 1995).
Bertsekas, D. Convex Optimization Theory, Vol. 1 (Athena Scientific, 2009).
Le Gallo, M. et al. Mixed-precision in-memory computing. Nat. Electron. 1, 246â253 (2018).
Pham, T.-A. et al. Three-dimensional optical diffraction tomography with Lippmann-Schwinger model. IEEE Trans. Comput. Imaging 6, 727â738 (2020).
Yurkin, M. A. & Hoekstra, A. G. The discrete dipole approximation: an overview and recent developments. J. Quant. Spectrosc. Radiat. Transf. 106, 558â589 (2007).
Molesky, S. et al. Inverse design in nanophotonics. Nat. Photonics 12, 659â670 (2018).
Calvo-Fullana, M., Paternain, S., Chamon, L. F. O. & Ribeiro, A. State augmented constrained reinforcement learning: Overcoming the limitations of learning with rewards. IEEE Trans. Autom. Control, 69, 4275â4290 (2024).
Kalogerias, D. S. & Petropulu, A. P. Matrix Completion in colocated MIMO radar: recoverability, bounds & theoretical guarantees. IEEE Trans. Signal Process. 62, 309â321 (2014).
Bao, F. et al. Heat-assisted detection and ranging. Nature 619, 743â748 (2023).
Horodynski, M., Kühmayer, M., Ferise, C., Rotter, S. & Davy, M. Anti-reflection structure for perfect transmission through complex media. Nature 607, 281â286 (2022).
Acknowledgements
The authors would like to thank Mario Junior Mencagli for useful discussions and preliminary experimental surveys on the subject. D.C.T. acknowledges Luiz F. O. Chamon and Juan Cerviño for the useful inputs and discussions regarding the constrained optimization algorithm. This work is supported in part by the Air Force Office of Scientific Research (AFOSR) Multidisciplinary University Research Initiative (MURI) grant numbers FA9550-17-1-0002 and FA9550-21-1-0312.
Author information
Authors and Affiliations
Contributions
N.E. conceived the idea for the reconfigurable device that solves equations, acquired the funds, and supervised the project. D.C.T. developed further the relevant theories and analyses of the project. B.E. designed and programmed the device and the deviceâs calibration routine. D.C.T. and B.E. assembled, built, and tested the components, and performed simulations and experimental measurements. D.C.T. and B.E. developed the numerical examples. All the authors discussed the results. D.C.T. wrote the first draft of the manuscript and D.C.T., B.E., and N.E. discussed, developed, and edited the final version of the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks the anonymous, reviewers for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisherâs note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the articleâs Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the articleâs Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Tzarouchis, D.C., Edwards, B. & Engheta, N. Programmable wave-based analog computing machine: a metastructure that designs metastructures. Nat Commun 16, 908 (2025). https://doi.org/10.1038/s41467-025-56019-1
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-56019-1