Abstract
A polarization-dependent silicon nano-antennas metagrating (PSNM) is proposed for parallel polarization transformation by engineering diffraction orders, upon which a compact Mueller microscopy system is implemented for subtissue-level polarization extraction. The polarization-dependent metagrating is designed using matrix Fourier optics and nonlinear optimization with four diffraction orders described by waveplate-like Jones matrices, which is encoded by nano-antennas combining geometric and propagation phases. The measured phase delay and orientation of each diffraction order of the metagrating deviate by less than 6.7 % from the design values, and the overall diffraction efficiency reaches 70.89 % with a coefficient of variation of 0.021. A transmissive PSNM Mueller microscopy system is developed by directly embedding the metagrating into an infinity-corrected microscopic optical path, which extracts subtissue-level polarization distributions of biological sections over a 152 μm × 152 μm field of view with reduced measurement redundancy, facilitating the differentiation and staging of pathological tissues for potential stain-free diagnostic applications.
1 Introduction
Polarization imaging reveals valuable microstructural details invisible to conventional microscopy, allowing observation of low-contrast, transparent biological tissues with anisotropic biomolecules [1], [2], [3], [4], [5]. Mueller matrix microscopy provides the most comprehensive microstructural information among polarization imaging techniques and is widely used in morphology and pathology [6], [7], [8], [9], especially for analyzing anisotropic changes in diseased tissues like cervical, skin, and other tumors [10], [11], [12], [13], [14], [15].
Optical metasurfaces offer precise modulation of light amplitude, phase, and polarization at subwavelength scales [16], [17], [18], [19], [20], [21], supporting applications such as versatile on-chip manipulation [22], [23], spin-controlled generation [24], full-color in situ imaging [25], chiral imaging [26], polarization holography [27], [28], [29], and Stokes imaging [30], [31]. Compact metasurface-based Stokes imaging has been proposed using interleaved or superpixel structures to spatially separate intensity information of different polarization states. Polarized subimages are spatially separated by interleaved metalenses, based on which full-Stokes imaging is reconstructed with an ultra-high numerical aperture (NA) [32]. By optimizing meta-atom lattice constant and scale of interleaved metalens, the imaging resolution, measured with a United States Air Force (USAF) test target, has reached 2.19 μm, with reduced electromagnetic crosstalk between interleaved sections. Orthogonal polarized components of incident image are selected and focused to corresponding pixel regions by superpixel structure, the intensities of which are detected superpixel by superpixel to reconstruct full Stokes imaging [33]. By trading spatial resolution, imaging polarimetry with a transmission efficiency of about 65 % is achieved using a superpixel configuration, overcoming the 50 % theoretical efficiency limit of traditional DOFP (division of focal plane) polarization cameras. Based on superpixel polarization imaging, metasurface-based Mueller microscopy is achieved by integrating ultra-thin multilayer nanograting filter arrays onto a CMOS sensor, providing exceptional compactness by eliminating redundant optical components and lengthy measurements typical of conventional systems, thereby facilitating advances in medical imaging and diagnostic science [34]. However, as the metasurface-based Mueller matrix microscopy is implemented using superpixels, each covering the area of six sensor pixels, the effective field of view is inherently restricted, resulting in reduced spatial resolution and increased vignetting [33], which in turn limits its applicability to subtissue-level imaging. Based on matrix Fourier optics [35], [36], [37], [38], a metagrating was designed using nanopillar arrays to control far-field polarization, achieving single-shot full-Stokes polarization camera for imaging distant objects without moving parts or specialized sensors [39], [40]. In this paper, a polarization-dependent silicon nano-antennas metagrating (PSNM) was proposed to achieve parallel polarization transformation without spatial multiplexing or a shared aperture, upon which a compact Mueller microscopy system was implemented for subtissue-level polarization imaging and anisotropy information extraction. Based on matrix Fourier optics, the metagrating is designed to preserve four diffraction orders with each possessing a Jones matrix in the form of a waveplate oriented in distinct directions. A nonlinear constrained optimization algorithm was employed to determine the phase profile of the designed 15 × 15 metagrating unit cell, achieving the desired Jones matrices in the retained diffraction orders with high and uniform diffraction efficiency. According to the optimized phase distribution, the metagrating structure was encoded by nano-antennas with both geometric and propagation phases. Then, the metagrating was fabricated on a fused silica substrate by EBL and ICP technique. Polarization and diffraction efficiencies of each fabricated metagrating order were experimentally measured, revealing maximum deviations of 6.7 % in phase delay and 4 % in fast axis orientation from the designed waveplates, with an overall diffraction efficiency of 70.89 % and a high uniformity quantified by a coefficient of variation of 0.021. By integrating the metagrating with a 100× objective lens, the transmissive PSNM Mueller microscopy system was constructed and eigenvalue-calibrated, featuring streamlined measurement procedures with a field of view (FOV) of 152 μm × 152 μm. Through Mueller matrix measurement conducted by the PSNM system and MMPD parameter extraction, the subtissue scale depolarization distribution of Epipremnum aureum leaf sections was measured, and the average depolarization was found to decrease with storage time, which is consistent with results obtained by conventional methods. Similar polarization analysis performed on mouse fibrotic liver and cervical cancer tissues identified total retardance as an effective metric for staging liver fibrosis and distinguishing cancerous from normal cervical tissue, highlighting the system’s potential application in portable, stain-free auxiliary cancer diagnosis and monitoring.
2 Design method
2.1 The mechanism for generating polarization response of diffraction orders
The structure schematic of polarization-dependent silicon nano-antennas metagrating (PSNM) is shown in Figure 1(a). The nano-antenna has a rectangular cross section with length l, width w, and orientation angle θ. The unit cell of PSNM generally includes p × q nano-antennas with an interelement separation d. The Jones matrix of nano-antenna at row a and column b within a unit cell is given as:
where:

PSNM structure construction. (a) Unitcell schematic of PSNM and magnified structure diagram of nano-antennas. (b) The framework of structural optimization algorithm for nano-antennas polarization-dependent metagrating. (c), (d) Swept propagation phase and amplitude transmission with x-polarized incidence, and (e), (f): with y-polarized incidence.
Illuminated by a normally incident uniform plane-wave
where
where
n and m are integers.
2.2 Optimization and design implementation
The phase profile and orientation distribution of nano-antennas metagrating is designed by a nonlinear constrained optimization with respect to a figure of merit. The polarization-dependent metagrating is expected to retain the four diffraction orders
The target orientation angles

where F is the merit function; superscript † represents the Hermitian conjugate, the trace
. g1 = 0 means the standard deviation of the diffraction efficiency among the retained diffraction orders is expected to be 0, to ensure that the “weights” of the Jones matrices
Based on Lagrange multipliers method, the optimization problem equation (7) is transformed into an unconstrained optimization problem:
where λ is the Lagrange multiplier vector,
The framework of nano-antennas polarization-dependent metagrating optimization is shown as Figure 1(b):
Step 1: Set the polarization-dependent metagrating variables
Step 2: Start with an initial value of optimization variable
Step 3: Perform the damped quasi-Newton algorithm to iterate the metagrating structure variables. In the k-th iteration, the gradient of the Lagrange function at current step is ∇L
k
, gradient increment is y
k−1 = ∇L
k
− ∇L
k−1, and last vector step length is s
k−1 = x
k
− x
k−1. Update the approximate Hesse matrix inverse
where I is the identity matrix. By the quasi-Newton searching equation:
The search direction d k is determined. Then Wolfe linear search principle is applied:
where
Step 4: Repeat Step 3, until ∇L
k
reaches the convergence precision ∇L
k
≤ ɛ, then output the total optimization variable
During the optimization process, only
The nano-antennas structure database was established by the numerical simulations, in which the amplitude transmission and propagation phase related to selected nano-antennas were also recorded. The height and lattice constant of α-Si nano-antennas were 710 nm and 550 nm, respectively. The length l and width w of each nano-antenna were swept in the range of 100 nm–320 nm with an interval of 2 nm at wavelength 808 nm. The propagation phase:
The nano-antennas structures of the polarization-dependent metagrating unit cell were encoded according to the optimized phase profile. The mean square error
At each spatial coordinate
3 Results
3.1 Metagrating characterization
3.1.1 Simulation
The polarization response of the four retained diffraction orders of the metagrating was simulated. The unit cell structure of metagrating was modeled in the Lumerical FDTD software with Periodical Boundary conditions under both normal incidence and oblique incidence at wavelength 808 nm. The polarization ellipse parameters of each retained diffraction order were monitored by a far-field polarization analysis group under linearly, circularly, and elliptically polarized incidence. The simulated polarization ellipse parameters, including azimuth angle α, ellipticity angle ɛ, and handedness δ, were converted to the unified Stokes parameters S 1, S 2, S 3:
where S 0 refers to the first element of the Stokes vector. The unified Stokes parameters S 1, S 2, S 3 of each retained diffraction order were simulated and displayed on the Poincaré sphere after being normalized by S 0 as shown in Figure 2(a)–(d), which are consistent with the theoretical waveplate response under both normal and oblique incidence. The diffraction efficiency of each order, defined as the mean relative diffraction efficiency under orthogonal incident polarizations, was also simulated as shown in Figure 2(e). The diffraction efficiencies of the four retained diffraction orders were relatively uniform with an average efficiency of 20.53 % and a coefficient of variation of 0.0121. The simulated operational bandwidth of the nano-antennas metagrating is 780–835 nm, and the details of which are shown in the Supplementary Section S3.

Simulation and characterization of PSNM: simulated/theoretical unified Stokes parameters of each retained diffraction orders under: (a) 0° linear polarized normal incidence, (b): elliptical polarized normal incidence, (c): right-circular polarized normal incidence, (d): 0° linear polarized; θ = 4°, φ = 45° oblique incidence. (e) The simulated diffraction efficiency of each order (m,n). (f) The experimental setup for the measurement of fabricated metagrating characterization. (g) The top view and magnified details of SEM images of the fabricated polarization-dependent silicon nano-antennas metagrating samples. (h) Theoretical, simulated, and experimental value of phase delay and fast axis orientation of each retained order. (i) Simulated and experimental diffraction efficiency of each retained diffraction order.
The Mueller matrix is employed to further characterize the polarization responses of each diffraction order of the metagrating. The Mueller matrix is a 4 × 4 matrix that describes the linear transformation between the incident and transmitted Stokes vectors, characterizing the intrinsic polarization properties of the sample. The Mueller matrix of each retained diffraction order was simulated, from which the phase delay and fast axis orientation of each order was calculated. Under the kth independent incident polarization states
where
By employing the Mueller Matrix Polar Decomposition (MMPD) method [44], polarization parameters such as depolarization, total retardance, linear phase retardance, and linear fast axis orientation are extracted from the Mueller matrix. The waveplate’s birefringent behavior of each retained diffraction order is characterized by two parameters: the phase delay, which corresponds to the linear phase retardance, and the fast axis orientation, which corresponds to the linear fast axis orientation. Both parameters were shown as the red bars in Figure 2(g). Among four retained diffraction orders, the maximum deviation of phase delay between the simulated values and designed wave plate parameters was 2 %, and that of fast axis orientation was 0.81 %.
3.1.2 Fabrication
The fused silica substrate was ultrasonically cleaned using acetone, isopropanol, and deionized water. A 710 nm thick α-Si film was deposited on the fused silica substrate by plasma-enhanced chemical vapor deposition (PECVD) with a 5 % mixture of silane in argon at 300 °C. A photoresist layer (ZEP 520A) with 170 nm thickness was spin-coated on the surface of α-Si film. The metagrating pattern was written in the photoresist layer via electro-beam lithography (EBL) at a beam current of 2 nA followed by development in a resist developer (ZED-N50, Zeon Chemicals). A chromium layer with 25 nm thickness was deposited on the patterned photoresist layer. Then the photoresist layer was lifted off and the metagrating pattern was transferred to the chromium layer as a hard mask. Finally, the single-layer α-Si metagrating sample was fabricated by inductively coupled plasma (ICP) etching technique and then cleaned with acetone, isopropanol, and deionized water.
3.1.3 Fabricated metagrating characterization
The Mueller matrix of each retained diffraction order of the fabricated metagrating was experimentally measured. The experimental setup for the measurement of fabricated metagrating characterization is shown in Figure 2(f). Independent incident polarization states
3.2 PSNM Mueller microscopy
3.2.1 PSNM Mueller microscopy setup and calibration measurement
The optical setup of the PSNM Mueller microscopy system was shown in the Figure 3(a). The metagrating was placed between the objective lens (HC FL PLAN 100×, Leica, FOV = 250 μm in diameter) and CMOS camera. The laser wavelength used in our experiments is 808 nm, which offers superior transmittance for biological tissues. The resolution of the PSNM Mueller microscopy system was 1.55 μm, which was measured through a USAF test target (HIGHRES-1, Newport). As shown in Figure 4(d), the image of the eighth group pattern of the resolution test target was measured by the Mueller microscopy system. The elements in the third row are clearly resolved, whereas those in the fourth row are only partially distinguishable. Based on the table in Figure 4(e), the line width of the third-row element in Group 8 is 1.55 μm; therefore, the spatial resolution of our system has reached at least 1.55 μm. The FOV of each diffraction channel was 152 μm × 152 μm effectively preserving the maximal inscribed square region within the objective lens’s original FOV.

Calibration and verification of PSNM Mueller microscopy system. (a) The optical setup of the PSNM Mueller microscopy system. LP1 & LP2, linear polarizer, LP2 was SM1-mounted at the CMOS camera entrance; QWP1, quarter-wave plate; OBJ, microscope objective lens; L1 & L2 & L3, convex lenses (f1 = f2 = 100 mm, f3 = 25 mm); AA, adjustable aperture; MS, the fabricated metagrating sample; CP1 and CP2, calibration plane; CMOS, complementary metal oxide semiconductor. (b) The experimental eigenvalues λ 1, λ 2 … … to λ 16 of K. (c) Normalized Mueller matrix image of the verification sample.

Resolution measurement and biological tissue characterization of the calibrated PSNM Mueller microscopy system. (a) Original intensity image captured by the COMS camera while the orientation angle of the quarter waveplate of the PSNM Mueller microscopy system was 75°, the bright central region originated from stray light, and the defocused image generated by zero-order light, which did not interact with the metagrating. (b) Depolarization distribution obtained by MMPD and the corresponding mean value of the fresh-cut tissue sections of Epipremnum aureum leaf measured at different time points. From left to right: 0 h, 3 h, and 6 h. (c) Total retardance distribution as well as the corresponding mean value for each stage of fibrotic tissue sections. From left to right: F0 (normal), F2 (moderate fibrosis), and F4 (Cirrhosis). (d) The image of the eighth group pattern of the resolution test target (HIGHRES-1, Newport). (e) Correspondence table of feature line width for elements in each group of the USAF resolution test target (HIGHRES-1, Newport).
The system was then calibrated and verified. The calibration samples were a quarter waveplate of which the preset fast axis orientation was about 30°, a linear polarizer (#OCZ200, BOCIC; theoretical transmission 78 %) of which the azimuth angle were preset to approximately 0° and 45°, respectively. Verification sample for PSNM Mueller microscopy system was a linear polarizer with preset azimuth angle about −45°. The samples, including the calibration samples and the verification sample, were individually positioned on the working platform of the microscope objective lens. The orientation angle of the quarter waveplate QWP1 of the PSNM Mueller microscopy system was changed (in our experiment were 50°, 75°, 105°, and 130°, respectively) to control the polarization incidence. The intensity distributions of four retained diffraction channels of the PSNM Mueller microscopy system were captured by the CMOS camera under different polarization incidence as
3.2.2 Data analysis
The Eigenvalue Calibration Method (ECM) [45] was adapted to calibrate the PSNM Mueller microscopy system. Based on the measured intensity projection matrices, the Sylvester equations of each calibration sample were summed:
where
The Mueller matrix distribution
which is shown in Figure 3(c). The maximal deviation of the mean Mueller matrix element of the verification sample compared with the theoretical one was as small as 0.0094, primarily attributed to the random noise in the captured light intensities. Therefore, the calibration of the PSNM Mueller microscopy system was completed and validated.
The air sample was also measured for verification based on a backward eigenvalue calibration. In the backward calibration, the sample is placed on the calibration plane CP2. The instrument matrix A is derived through a procedure similar to that used for the modulation matrix W, rather than directly from the null response. The Mueller matrix of the air sample was determined using equation (16) by employing the modulation matrix W, the backward calibration-determined A, and the null response D 0. The maximal deviation of the mean Mueller matrix element of the air sample compared with the theoretical value was 0.0143, which is comparable to previously reported high performance system [46], [47]. The details of air sample verification are shown in Supplementary Section S2.
3.2.3 Polarization characteristics of biological tissue explored by calibrated PSNM Mueller microscopy system
The calibrated PSNM Mueller microscopy system was used to investigate the subtissue-level polarization distributions of various biological samples.
The fresh-cut tissue sections of E. aureum leaf were placed at the working plane position in the experimental setup shown as Figure 3(a). The intensity projection matrix distribution was measured using the calibrated PSNM Mueller microscopy system at three distinct time points: immediately after sample preparation (0 h), and subsequently at 3-h and 6-h intervals, following the same protocol as outlined in Section 3.2.1. Then the Mueller matrix distribution of E. aureum leaf section was calculated from equation (16) as shown in Supplementary Figure S5. Based on the MMPD method mentioned in Section 3.1.1, the measured Mueller matrix distributions were analyzed to extract polarization parameters at each time point, among which the depolarization parameter exhibited a pronounced change, as shown in Figure 4(b). The average depolarization decreased over storage time due to variations in the water content of plant leaves, a trend which is consistent with results obtained from conventional Mueller microscopy [48].
The polarization characteristics of fibrotic tissue sections from mouse liver were subsequently measured by the calibrated PSNM Mueller microscopy system. The Mueller matrix distribution for normal stage F0, moderate liver fibrosis stage F2, and cirrhosis stage F4 of the unstained tissue paraffin sections were measured as shown in the Supplementary Figure S6. Among the polarization parameters extracted by the MMPD method mentioned in Section 3.1.1, the total retardance parameter exhibited substantial changes across different stages of fibrosis, as shown in Figure 4(c). The progression of tissue fibrosis stages led to an increase in the mean of retardance distribution; therefore, the total retardance parameter is considered as a potential polarization-based metric for characterizing the stages of liver fibrosis.
Mueller matrix measurements and MMPD parameter extraction were also conducted using the calibrated PSNM Mueller microscopy system to characterize the polarization properties of mouse cervical cancer tissue. Comparative polarization analyses were performed on parallel experimental groups, each containing one cancerous tissue section and one normal tissue section. Among the polarization parameters extracted from the Mueller matrix distributions of each parallel experimental group, the total retardance was identified as a highly effective parameter for distinguishing between cancerous and normal tissue sections as shown in the Figure 5(a). Normal cervical tissue groups were characterized by a considerably higher mean total retardance than cancerous tissue groups as shown in the Figure 5(b), since tumor-associated fibrosis was a common phenomenon during cancer progression [49], [50], [51]. The calibrated PSNM Mueller microscopy system facilitated detailed polarization analysis from tissue sections, holding promise in cancer diagnosis.

Comparative polarization analyses between cancerous and normal cervical tissue sections. (a) The total retardance distribution of each parallel experiment group. The first column represented cancerous cervical tissues from each group, while the second column represented normal tissues. (b) The mean of total retardance of normal and cancerous cervical tissues from each parallel experimental group.
4 Discussion
The single-layer metagrating achieved microscopic imaging with separated polarization control via engineered waveplate responses at designated diffraction orders without shared aperture or spatial multiplexing, upon which a compact Mueller microscopy system was facilitated. Despite fabrication imperfections in the metagrating, the PSNM Mueller microscopy system maintained high measurement precision after appropriate calibration and reduced the measurement requirements due to the parallel design. The experiment was conducted at the wavelength of 808 nm, within a spectral region recognized for its effective penetration in biological tissue. By choosing a waveplate as the target Jones matrix for each diffraction order, more optical energy can be transmitted into the imaging system, which is particularly advantageous for detecting weak transmitted images from biological tissues. Moreover, the Jones matrix of each diffraction order of the metagrating can also be engineered to emulate the responses of various polarization elements, which in turn enables the realization of integrated and parallel polarization imaging systems designed for specific functionalities, such as polarization remote sensing [52] and robot perception [53].
In conclusion, a polarization-dependent silicon nano-antennas metagrating (PSNM) was proposed for parallel polarization transformation, facilitating the development of a compact PSNM Mueller microscopy system with reduced measurement requirements. Based on matrix Fourier optics, the metagrating was engineered to yield quarter-waveplate-like Jones matrices with distinct orientations in the retained diffraction orders. Nonlinear constrained optimization was applied to ensure uniformity and high overall efficiency across these orders. Experimental characterization of the fabricated metagrating revealed maximum deviations of 6.7 % in phase delay and 4 % in fast axis orientation from the designed waveplates, with an overall diffraction efficiency of 70.89 % and a high uniformity quantified by a coefficient of variation of 0.021. The metagrating was employed to implement a compact Mueller matrix microscopy system with a 152 μm × 152 μm FOV and 1.55 μm resolution. The maximal deviation between the experimentally averaged and theoretical Mueller matrix elements of the verification sample was as low as 0.0094. The PSNM-based Mueller matrix measurements and MMPD parameter extraction conducted on E. aureum leaf sections have revealed a storage-time-dependent decrease in average depolarization consistent with conventional results. Similar polarization analysis performed on mouse fibrotic liver and cervical cancer tissues has identified total retardance as a promising metric for staging liver fibrosis and distinguishing cancerous from normal cervical tissue, highlighting the system’s potential for application in stain-free cancer diagnosis.
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 61927822
-
Research funding: This work was supported by the National Natural Science Foundation of China (NSFC; 61927822).
-
Author contributions: QL and XD proposed the original idea and performed the theoretical analysis. QL coded the framework of the method. ZL contributed to developing the measurement method. QL, JL, and GC performed the experiments and analyzed the results. DL contributed to preparing experimental sections. QL and XD wrote the manuscript and revised it. The project was supervised by XD. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript.
-
Conflict of interest: Authors state no conflicts of interest.
-
Data availability: Data supporting this study are available from the corresponding author on reasonable request.
References
[1] N. Ghosh and A. I. Vitkin, “Tissue polarimetry: Concepts, challenges, applications, and outlook,” J. Biomed. Opt., vol. 16, no. 11, 2011, Art. no. 110801, https://doi.org/10.1117/1.3652896.Search in Google Scholar PubMed
[2] W. Groner et al.., “Orthogonal polarization spectral imaging: A new method for study of the microcirculation,” Nat. Med., vol. 5, no. 10, pp. 1209–1212, 1999. https://doi.org/10.1038/13529.Search in Google Scholar PubMed
[3] R. Turcotte, J. M. Mattson, J. W. Wu, Y. Zhang, and C. P. Lin, “Molecular order of arterial collagen using circular polarization second-harmonic generation imaging,” Biophys. J., vol. 110, no. 3, pp. 530–533, 2016, https://doi.org/10.1016/j.bpj.2015.12.030.Search in Google Scholar PubMed PubMed Central
[4] W. Mickols, M. Maestre, and I. Tinoco, “Differential polarization microscopy of changes in structure in spermatocyte nuclei,” Nature, vol. 328, no. 6129, pp. 452–454, 1987. https://doi.org/10.1038/328452a0.Search in Google Scholar PubMed
[5] J. Bailey et al.., “Circular polarization in Star- formation regions: Implications for biomolecular homochirality,” Science, vol. 281, no. 5377, pp. 672–674, 1998. https://doi.org/10.1126/science.281.5377.672.Search in Google Scholar
[6] S. Badieyan et al.., “Detection and discrimination of bacterial colonies with mueller matrix imaging,” Sci. Rep., vol. 8, no.1, 2018, Art. no. 10815, https://doi.org/10.1038/s41598-018-29059-5.Search in Google Scholar PubMed PubMed Central
[7] J. Qi and D. S. Elson, “A high definition Mueller polarimetric endoscope for tissue characterisation,” Sci. Rep., vol. 6, no.1, 2016, Art. no. srep25953, https://doi.org/10.1038/srep25953.Search in Google Scholar PubMed PubMed Central
[8] F. Fanjul-Vélez, N. Ortega-Quijano, and J. L. Arce-Diego, “Polarimetry group theory analysis in biological tissue phantoms by Mueller coherency matrix,” Opt. Commun., vol. 283, no. 22, pp. 4525–4530, 2010, https://doi.org/10.1016/j.optcom.2010.04.074.Search in Google Scholar
[9] S. Alali and A. Vitkin, “Polarized light imaging in biomedicine: Emerging Mueller matrix methodologies for bulk tissue assessment,” J. Biomed. Opt., vol. 20, no. 6, p. 061104, 2015. https://doi.org/10.1117/1.jbo.20.6.061104.Search in Google Scholar PubMed
[10] Y. Wang et al.., “Differentiating characteristic microstructural features of cancerous tissues using Mueller matrix microscope,” Micron, vol. 79, pp. 8–15, 2015, https://doi.org/10.1016/j.micron.2015.07.014.Search in Google Scholar PubMed
[11] A. Pierangelo et al.., “Polarimetric imaging of uterine cervix: A case study,” Opt. Express, vol. 21, no. 12, pp. 14120–14130, 2013, https://doi.org/10.1364/oe.21.014120.Search in Google Scholar
[12] D. Ivanov et al.., “Polarization and depolarization metrics as optical markers in support to histopathology of ex vivo colon tissue,” Biomed. Opt. Express, vol. 12, no. 7, pp. 4560–4572, 2021. https://doi.org/10.1364/boe.426713.Search in Google Scholar PubMed PubMed Central
[13] I. Ahmad, A. Khaliq, M. Iqbal, and S. Khan, “Mueller matrix polarimetry for characterization of skin tissue samples: A review,” Photodiagn. Photodyn. Ther., vol. 30, 2020, Art. no. 101708, https://doi.org/10.1016/j.pdpdt.2020.101708.Search in Google Scholar PubMed
[14] J. C. Setiadi, J. Bonaventura, T. G. Knapp, S. Duan, J. L. Merchant, and T. W. Sawyer, “Mueller matrix polarization imaging of gastrinoma shows promise for tumor localization,” in Proc. SPIE 12391, Label-Free Biomedical Imaging and Sensing (LBIS) 2023, 2023, p. 123910E.10.1117/12.2649407Search in Google Scholar
[15] V. A. Ushenko et al.., “3D Mueller matrix mapping of layered distributions of depolarisation degree for analysis of prostate adenoma and carcinoma diffuse tissues,” Sci. Rep., vol. 11, no. 1, p. 5162, 2021. https://doi.org/10.1038/s41598-021-83986-4.Search in Google Scholar PubMed PubMed Central
[16] S. Wang et al.., “A broadband achromatic metalens in the visible,” Nat. Nanotechnol., vol. 13, no. 3, pp. 227–232, 2018. https://doi.org/10.1038/s41565-017-0052-4.Search in Google Scholar PubMed
[17] F. Yue et al.., “Nonlinear Mid-Infrared metasurface based on a phase-change material,” Laser Photonics Rev., vol. 15, no. 3, 2021, Art. no. 2000373, https://doi.org/10.1002/lpor.202000373.Search in Google Scholar
[18] T. Zheng, Y. Gu, H. Kwon, G. Roberts, and A. Faraon, “Dynamic light manipulation via silicon-organic slot metasurfaces,” Nat. Commun., vol. 15, no. 1, p. 1557, 2024. https://doi.org/10.1038/s41467-024-45544-0.Search in Google Scholar PubMed PubMed Central
[19] M. Deng et al.., “Dielectric metasurfaces for broadband phase-contrast relief-like imaging,” Nano Lett., vol. 24, no. 46, pp. 14641–14647, 2024, https://doi.org/10.1021/acs.nanolett.4c03695.Search in Google Scholar PubMed
[20] Z. Xi et al.., “Room-temperature mid-infrared detection using metasurface-absorber-integrated phononic crystal oscillator,” Laser Photonics Rev., 2025, vol. 19, no. 20, Art. no. e00498, https://doi.org/10.1002/lpor.202500498.Search in Google Scholar
[21] A. Basiri, M. Z. E. Rafique, J. Bai, S. Choi, and Y. Yao, “Ultrafast low-pump fluence all-optical modulation based on graphene-metal hybrid metasurfaces,” Light: Sci. Appl., vol. 11, no. 1, p. 102, 2022. https://doi.org/10.1038/s41377-022-00787-8.Search in Google Scholar PubMed PubMed Central
[22] F. Ding, B. Chang, Q. Wei, L. Huang, X. Guan, and S. I. Bozhevolnyi, “Versatile polarization generation and manipulation using dielectric metasurfaces,” Laser Photonics Rev., vol. 14, no. 11, 2020, Art. no. 2000116, https://doi.org/10.1002/lpor.202000116.Search in Google Scholar
[23] S. Sande et al.., “On-chip emitter-coupled meta-optics for versatile photon sources,” Phys. Rev. Lett., vol. 135, no. 8, 2025, Art. no. 086902, https://doi.org/10.1103/klq1-wjjg.Search in Google Scholar PubMed
[24] S. im Sande, Y. Deng, S. I. Bozhevolnyi, and F. Ding, “Spin-controlled generation of a complete polarization set with randomly-interleaved plasmonic metasurfaces,” Opto-Electron. Adv., vol. 7, no. 8, 2024, Art. no. 240076, https://doi.org/10.29026/oea.2024.240076.Search in Google Scholar
[25] M. Garcia, C. Edmiston, R. Marinov, A. Vail, and V. Gruev, “Bio-inspired color-polarization imager for real-time in situ imaging,” Optica, vol. 4, no. 10, pp. 1263–1271, 2017. https://doi.org/10.1364/optica.4.001263.Search in Google Scholar
[26] M. Khorasaninejad et al.., “Multispectral chiral imaging with a metalens,” Nano Lett., vol. 16, no. 7, pp. 4595–4600, 2016, https://doi.org/10.1021/acs.nanolett.6b01897.Search in Google Scholar PubMed
[27] D. Wang et al.., “Decimeter-depth and polarization addressable color 3D meta-holography,” Nat. Commun., vol. 15, no. 1, p. 8242, 2024. https://doi.org/10.1038/s41467-024-52267-9.Search in Google Scholar PubMed PubMed Central
[28] H. Zhou et al.., “Polarization-encrypted orbital angular momentum multiplexed metasurface holography,” ACS Nano, vol. 14, no. 5, pp. 5553–5559, 2020, https://doi.org/10.1021/acsnano.9b09814.Search in Google Scholar PubMed PubMed Central
[29] R. Zhao et al.., “Encoding arbitrary phase profiles to 2D diffraction orders with controllable polarization states,” Nanophotonics, vol. 12, no. 1, pp. 155–163, 2023, https://doi.org/10.1515/nanoph-2022-0707.Search in Google Scholar PubMed PubMed Central
[30] G. Soma, K. Komatsu, C. Ren, Y. Nakano, and T. Tanemura, “Metasurface-enabled non-orthogonal four-output polarization splitter for non-redundant full-stokes imaging,” Opt. Express, vol. 32, no. 20, pp. 34207–34222, 2024. https://doi.org/10.1364/oe.529389.Search in Google Scholar PubMed
[31] Y. Zhang et al.., “Crosstalk-free achromatic full Stokes imaging polarimetry metasurface enabled by polarization-dependent phase optimization,” Opto-Electron. Adv., vol. 5, no. 11, 2022, Art. no. 220058, https://doi.org/10.29026/oea.2022.220058.Search in Google Scholar
[32] Z. Huang et al.., “High-resolution metalens imaging polarimetry,” Nano Lett., vol. 23, no. 23, pp. 10991–10997, 2023, https://doi.org/10.1021/acs.nanolett.3c03258.Search in Google Scholar PubMed
[33] E. Arbabi, S. Mahsa Kamali, A. Arbabi, and A. Faraon, “Full-Stokes imaging polarimetry using dielectric metasurfaces,” ACS Photonics, vol. 5, no. 8, pp. 3132–3140, 2018, https://doi.org/10.1021/acsphotonics.8b00362.Search in Google Scholar
[34] J. Zuo et al.., “Metasurface-based Mueller matrix microscope,” Adv. Funct. Mater., vol. 34, no.45, 2024, Art. no. 2405412, https://doi.org/10.1002/adfm.202405412.Search in Google Scholar
[35] A. Zaidi, N. A. Rubin, A. H. Dorrah, J.-S. Park, and F. Capasso, “Generalized polarization transformations with metasurfaces,” Opt. Express, vol. 29, no. 24, pp. 39065–39078, 2021. https://doi.org/10.1364/oe.442844.Search in Google Scholar PubMed
[36] I. Moreno, M. J. Yzuel, J. Campos, and A. Vargas, “Jones matrix treatment for polarization Fourier optics,” J. Mod. Opt., vol. 51, no. 14, pp. 2031–2038, 2004. https://doi.org/10.1080/09500340408232511.Search in Google Scholar
[37] I. Moreno, C. Iemmi, J. Campos, and M. J. Yzuel, “Jones matrix treatment for optical Fourier processors with structured polarization,” Opt. Express, vol. 19, no. 5, pp. 4583–4594, 2011. https://doi.org/10.1364/oe.19.004583.Search in Google Scholar PubMed
[38] N. A. Rubin, A. Zaidi, A. H. Dorrah, Z. Shi, and F. Capasso, “Jones matrix holography with metasurfaces,” Sci. Adv., vol. 7, no. 33, 2021, Art. no. eabg7488, https://doi.org/10.1126/sciadv.abg7488.Search in Google Scholar PubMed PubMed Central
[39] N. A. Rubin, G. D’Aversa, P. Chevalier, Z. Shi, W. T. Chen, and F. Capasso, “Matrix Fourier optics enables a compact full-stokes polarization camera,” Science, vol. 365, no. 6448, 2019, Art. no. eaax1839, https://doi.org/10.1126/science.aax1839.Search in Google Scholar PubMed
[40] N. A. Rubin, P. Chevalier, M. Juhl, M. Tamagnone, R. Chipman, and F. Capasso, “Imaging polarimetry through metasurface polarization gratings,” Opt. Express, vol. 30, no. 6, pp. 9389–9412, 2022. https://doi.org/10.1364/oe.450941.Search in Google Scholar
[41] M. Losurdo and K. Hingerl, Eds. Ellipsometry at the Nanoscale, Berlin, Heidelberg, Springer, 2013.10.1007/978-3-642-33956-1Search in Google Scholar
[42] J. Scott Tyo, “Design of optimal polarimeters: Maximization of signal-to-noise ratio and minimization of systematic error,” Appl. Opt., vol. 41, no. 4, pp. 619–630, 2002. https://doi.org/10.1364/ao.41.000619.Search in Google Scholar PubMed
[43] C. Xu and J. Zhang, “A survey of quasi-newton equations and quasi-newton methods for optimization,” Ann. Oper. Res., vol. 103, no. 1, pp. 213–234, 2001. Available at: https://doi.org/10.1023/a:1012959223138.10.1023/A:1012959223138Search in Google Scholar
[44] S.-Y. Lu and R. A. Chipman, “Interpretation of Mueller matrices based on polar decomposition,” J. Opt. Soc. Am. A, vol. 13, no. 5, pp. 1106–1113, 1996, https://doi.org/10.1364/josaa.13.001106.Search in Google Scholar
[45] E. Compain, S. Poirier, and B. Drevillon, “General and self-consistent method for the calibration of polarization modulators, polarimeters, and Mueller-matrix ellipsometers,” Appl. Opt., vol. 38, no. 16, pp. 3490–3502, 1999. https://doi.org/10.1364/ao.38.003490.Search in Google Scholar PubMed
[46] Y. Li et al.., “Holistic and efficient calibration method for Mueller matrix imaging polarimeter with a high numerical aperture,” Appl. Opt., vol. 61, no. 33, pp. 9937–9945, 2022. https://doi.org/10.1364/ao.474531.Search in Google Scholar
[47] S. Sheng et al.., “Eigenvalue calibration method for dual rotating-compensator Mueller matrix polarimetry,” Opt. Lett., vol. 46, no. 18, pp. 4618–4621, 2021. https://doi.org/10.1364/ol.437542.Search in Google Scholar
[48] A. Van Eeckhout et al.., “Depolarizing metrics for plant samples imaging,” PLoS One, vol. 14, no. 3, 2019, Art. no. e0213909, https://doi.org/10.1371/journal.pone.0213909.Search in Google Scholar PubMed PubMed Central
[49] H. Fujimoto et al.., “Tumor-associated fibrosis: A unique mechanism promoting ovarian cancer metastasis and peritoneal dissemination,” Cancer Metastasis Rev., vol. 43, no. 3, pp. 1037–1053, 2024. https://doi.org/10.1007/s10555-024-10169-8.Search in Google Scholar PubMed PubMed Central
[50] H. Chen et al.., “Decoding tumor-fibrosis interplay: Mechanisms, impact on progression, and innovative therapeutic strategies,” Front. Pharmacol., vol. 15, 2024, Art. no. 1491400, https://doi.org/10.3389/fphar.2024.1491400.Search in Google Scholar PubMed PubMed Central
[51] A. Naik and A. Leask, “Tumor-associated fibrosis impairs the response to immunotherapy,” Matrix Biol., vol. 119, pp. 125–140, 2023, https://doi.org/10.1016/j.matbio.2023.04.002.Search in Google Scholar PubMed
[52] J. Scott Tyo, D. L. Goldstein, D. B. Chenault, and J. A. Shaw, “Review of passive imaging polarimetry for remote sensing applications,” Appl. Opt., vol. 45, no. 22, pp. 5453–5469, 2006. https://doi.org/10.1364/ao.45.005453.Search in Google Scholar PubMed
[53] C. Taglione, C. Mateo, and C. Stolz, “Polarimetric imaging for robot perception: A review,” Sensors, vol. 24, no. 14, p. 4440, 2024. https://doi.org/10.3390/s24144440.Search in Google Scholar PubMed PubMed Central
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2025-0405).
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Articles in the same Issue
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry