Abstract
Berry curvature that describes local geometrical properties of energy bands can elucidate many fascinating phenomena in solid-state, photonic, and phononic systems, given its connection to global topological invariants such as the Chern number. Despite its significance, the observation of Berry curvature poses a substantial challenge since wavefunctions are deeply embedded within the system. Here, we theoretically propose a correspondence between the geometry of far-field polarization and the underneath band topology in non-Hermitian systems, thus providing a general method to fully capture the Berry curvature without strongly disturbing the eigenstates. We further experimentally observe the Berry curvature in a honeycomb photonic crystal slab from polarimetry measurements and quantitatively obtain the nontrivial valley Chern number. Our work reveals the feasibility of retrieving the bulk band topology from escaping photons and paves the way to exploring intriguing topological landscapes in non-Hermitian systems.
Similar content being viewed by others
Introduction
Topology, namely the mathematics of conserved properties under continuous deformations, is creating a range of new opportunities throughout matters, photonics, phononics, and other wave systems1,2,3,4,5,6. To characterize the topology in physics, the Berry curvature7,8 is an essential concept that describes the local, gauge-invariant, geometric manifestation of the wavefunctions in the parameter space, which is closely related to the global topological invariants such as a variety of Chern numbers9,10,11,12,13. However, since Berry curvature belongs to the intrinsic topological property of wavefunctions, it is usually deeply bound inside the system and difficult to observe. Although using tomography can reconstruct the wavefunctions in some particular scenarios14,15, much effort has been devoted to retrieving the Berry curvature from its external consequence in physics. Examples include Hall drift in driving optical lattice16,17,18 or synthetic gauge field19,20, Aharonov–Bohm interference of magnetic-controlled ultracold atom21,22,23,24, and pseudospin25,26,27,28,29,30 or dichroism31,32 in exciton-polariton-correlated material. Even though the specific physics varies, the mentioned observations of Berry curvature generally rely on the strong light-matter interaction to imprint the topological features of bulk wavefunction to external observables, and thus they should be categorized into the class of “strong measurement”33, namely the observation strongly interfering with the system. In comparison, the method of measuring the Berry curvature without much disturbing the eigenstates34 remains absent.
It is well-recognized that open photonic systems necessarily lose photons to their surroundings. As a major route of energy exchange, the far-field radiation naturally carries information about wavefunctions and thus allows direct access to the intrinsic bulk topology, which would be conventionally thought impossible. Here, the escaping photons simply act as the “messengers” that weakly interact with the system, and could bridge the bulk band and radiation. For example, some advances in probing Chern numbers from the far field are reported, in which either phases and intensity35, or frequency and quality factor36 are chosen as observables. However, due to the difficulty in experiments and the lack of degree-of-freedom (DOF), they are probably not ideal observables to enable the direct observation of Berry curvature in general non-Hermitian multi-level photonic systems6,37,38,39,40. Hereby we propose that, the far-field polarization is a more applicable candidate because it provides sufficient DOFs, and more importantly, the orientation and ellipticity of polarization can be readily and robustly measured in a time-averaging manner. In fact, the nontrivial geometric features of far-field polarization41,42,43,44,45,46,47 has attracted much attention in recent years because they give rise to interesting physical consequences such as polarization half-charge around paired exceptional point48,49, vortex beam50, chiral devices with circular dichroism51,52, bound states in the continuum (BICs)53,54,55,56,57,58,59,60,61 and unidirectional guided resonances (UGRs)62,63,64. However, the intrinsic connection between the polarization and the bulk wavefunction is still unrevealed, and how to retrieve the Berry curvature from the far-field polarization features still remains an elusive question.
Here we theoretically establish a correspondence between the band topology and radiation geometry to experimentally observe the Berry curvature by characterizing the polarization of escaping photons from a non-Hermitian photonic crystal (PhC) slab system. We prove that, Berry curvatures of bulk bands can be captured by using observable bi-orthogonal polarization vectors in a non-Hermitian system40. As long as the radiation fields are smooth and well-defined, a full tomography of the Berry curvature can be realized by simply measuring the polarization of a number of radiation channels. Specifically, for a two-level system, only one radiation channel is sufficient. Accordingly, we experimentally observe the nontrivial Berry curvatures originating from the diabolic point (DP) in a honeycomb-latticed PhC slab by using a standard polarimetry measurement65. The Berry phases2,7,66 of γ ~ ±π are obtained in an individual valley by employing integrals of the observed Berry curvature, showing nontrivial valley-Chern numbers11,13 of \({C}_{v}^{{{{\mathcal{K}}}}} \sim \pm 1/2\) as expected, and thus quantitatively validate our method.
Results
The bulk-radiation correspondence of the Berry curvature
We start from an infinite two-dimensional (2D) PhC slab operating in the radiation continuum as schematically illustrated in Fig. 1A, which has translational symmetry in both x and y directions. According to the Bloch theorem, the system is ruled by a bulk momentum-space Hamiltonian \(\hat{H}({{{{\bf{k}}}}}_{\parallel })\), supporting multiple bulk Bloch modes whose cell-periodic Bloch function is denoted as \(| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\). The subscript n labels the energy bands, and k∥ is the in-plane Bloch wavevector. Due to the truncation in the z-direction, photons can partially escape from the PhC slab, radiating towards some specific out-of-plane directions owing to the diffraction of periodically modulated permittivity. The far-field radiation in each radiation channel Cq can be formulated by a vector of \(| {\Psi }_{n,q}\left.\right\rangle={[{c}_{x;n,q},{c}_{y;n,q}]}^{T}\), where cx,y;n,q are complex-valued electric field components of radiative wave in the x and y directions, respectively, and subscript q labels the radiation channel. The field components in s and p directions are discussed in Supplementary Materials section 1.
A The schematic of radiation process from the PhC slab to the far field in real space. The wavefunction of the Bloch mode \(| {u}_{n}\left.\right\rangle\) in the PhC slab is diffracted by the periodic lattice into several diffraction directions, acting as the radiation channels C1−3. Each channel (i.e. C3) depends on a 3D wavevector (i.e., k3), and can be described by the polarization vector (i.e., \(| {\Psi }_{n,3}\left.\right\rangle\)) marked as the spiral arrows. The 3D wavevector for each channel (i.e., k1 for channel C1) can be decomposed to the vertical component (i.e., kz,1) and in-plane component (i.e., β1), both determined by the Bloch wavevector k∥. B The “bulk-radiation correspondence" of Berry curvature in momentum space. The radiation polarization field (middle panel) can bridge the bulk Berry curvature Bn defined in wavefunction \(| {u}_{n}\left.\right\rangle\) (bottom panel) with the radiation Berry curvature \({B}_{n}^{r}\) defined in far-field radiation vector \(| {\Psi }_{n}\left.\right\rangle\) (top panel). cx,y are complex amplitudes of the radiative waves in the x−y plane.
Generally, as illustrated in Fig. 1A, the radiation vector is characterized by a three-dimensional (3D) wavevector kq, which can be decomposed into the in-plane component βq and the vertical component \({k}_{z,q}=\sqrt{{\omega }_{n}^{2}/{c}^{2}-| {{{\beta }}}_{q}{| }^{2}}\). Since the radiative waves are diffracted from the Bloch mode, βq, as well as kq, is directly determined from the Bloch wavevector k∥. To investigate the connection between bulk Bloch modes and its radiation, we focus on the dependence of radiative waves on the Bloch wavevector k∥ rather than the 3D wavevector kq. Consequently, at each k∥, the radiation process in a given channel Cq can be understood as a linear mapping of \({{{{\mathcal{P}}}}}_{q}:| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle \mapsto | {\Psi }_{n,q}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle={\hat{P}}_{q}({{{{\bf{k}}}}}_{\parallel })| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\), governed by a projection matrix denoted as \({\hat{P}}_{q}\). The detailed expressions of βq with respect to k∥ for each channel can be found in Supplementary Materials section 1. In the following discussions, we focus on a single channel, thus omitting the subscript q for simplicity.
The PhC slab we considered here is intrinsically non-Hermitian, due to the existence of radiation losses. Thus, the Berry curvature of the bulk Bloch bands in the 2D Brillouin zone (BZ) can be calculated from the wavefunctions as \({B}_{n}({{{{\bf{k}}}}}_{\parallel })=i{\nabla }_{{{{{\bf{k}}}}}_{\parallel }}\times \langle {v}_{n}({{{{\bf{k}}}}}_{\parallel })| {\nabla }_{{{{{\bf{k}}}}}_{\parallel }}{u}_{n}({{{{\bf{k}}}}}_{\parallel })\rangle\), denoted as bulk Berry curvature (bottom panel, Fig. 1B). Here \(\left\langle \right.{v}_{n}({{{{\bf{k}}}}}_{\parallel })|\) represents the left vector accompanied by the right vector \(| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\), ruled by the bi-orthogonal normalization relation of 〈vm(k∥)∣un(k∥)〉 = δmn. Accordingly, the right vector \(| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\) can be projected to the radiation vector \(| {\Psi }_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\) through the projection matrix \(\hat{P}\). If \(\hat{P}\) is an invertible matrix, the left vector \(\left\langle \right.{v}_{n}({{{{\bf{k}}}}}_{\parallel })|\) also corresponds to a radiation vector in the term of \(\left\langle \right.{\Phi }_{n}({{{{\bf{k}}}}}_{\parallel })|=\left\langle \right.{v}_{n}({{{{\bf{k}}}}}_{\parallel })| {\hat{P}^{-1}}({{{{\bf{k}}}}}_{\parallel })\). It is ready to derive that \(| {\Psi }_{n}\left.\right\rangle\) and \(\left\langle \right.{\Phi }_{n}|\) form another bi-orthogonal basis for the Hamiltonian of \({\hat{H}^{r}}=\hat{P}\hat{H}{\hat{P}^{-1}}\), leading to another Berry curvature \({B}_{n}^{r}({{{{\bf{k}}}}}_{\parallel })=i{\nabla }_{{{{{\bf{k}}}}}_{\parallel }}\times \langle {\Phi }_{n}({{{{\bf{k}}}}}_{\parallel })| {\nabla }_{{{{{\bf{k}}}}}_{\parallel }}{\Psi }_{n}({{{{\bf{k}}}}}_{\parallel })\rangle\) denoted as the “radiation Berry curvature” (top panel, Fig. 1B). We find that \({B}_{n}^{r}\) and Bn are related as \({B}_{n}^{r}({{{{\bf{k}}}}}_{\parallel })={B}_{n}({{{{\bf{k}}}}}_{\parallel })+i{\nabla }_{{{{{\bf{k}}}}}_{\parallel }}\times \left\langle \right.{v}_{n}({{{{\bf{k}}}}}_{\parallel })| {\hat{P}^{-1}}{\nabla }_{{{{{\bf{k}}}}}_{\parallel }}\hat{P}| {u}_{n}({{{{\bf{k}}}}}_{\parallel })\left.\right\rangle\). We further prove that (see Supplementary Materials section 1 for details), in the case that the matrix \(\hat{P}\) is smooth and doesn’t give rise to extra amplitude vortexes such as BICs that carry zero radiation (〈Ψn(k∥)∣Ψn(k∥)〉 = 0), the second term of \({B}_{n}^{r}\) vanishes, and the system follows a simple correspondence between the band topology and radiation topology, given by:
The above equation reveals that the escaping photons project the bulk Berry curvature onto the far-field radiation through the matrix \(\hat{P}\) (mid panel, Fig. 1B). Although the wavefunctions \(| {u}_{n}\left.\right\rangle\) and \(\left\langle \right.{v}_{n}|\) are difficult to access because they belong to the near-field features of bulk Bloch modes, the radiation vectors \(| {\Psi }_{n}\left.\right\rangle\) and \(\left\langle \right.{\Phi }_{n}|\) are directly observable. Therefore, Eq. (1) offers a possibility to extract the bulk band topology directly from the far-field radiation. The detailed derivation of such a bulk-radiation correspondence can be found in Supplementary Materials section 1. An example that poor-defined \(\hat{P}\) collapses the bulk-radiation correspondence is discussed in Supplementary Materials section 2. The physical origin and measurement methods of the left radiation vector \(\left\langle \right.{\Phi }_{n}|\) is discussed in the Supplementary Materials section 3.
In theory, the radiation in a particular radiation channel gives a perspective projection of the wavefunctions, and thus it is noteworthy to discuss whether the projection is complete. For a general vector system where N eigen modes exist, the number of internal DOFs of the system is N and both \(| {u}_{n}\left.\right\rangle\) and \(\left\langle \right.{v}_{n}|\) are N-dimensional vectors. Here we consider the case of N ≥ 2 since N = 1 refers to a scalar single-valuable system and thus only one scalar observable can fully characterize the system. As we stated above, for each bulk Bloch mode, one radiation channel can only characterize two DOFs (\({[{c}_{x;n},{c}_{y;n}]}^{T}\)), so \(\hat{P}\) should be a 2 × N matrix. For the case with N > 2, \(\hat{P}\) can’t be directly inverted, indicating that a single radiation channel can’t provide enough information to depict the bulk wavefunctions, and it’s necessary to measure multiple radiation channels simultaneously—observe the same object from different views. As a specific case, if the system is a simple two-level one with N = 2, the measurement of only one radiation channel is sufficient to make the projection of wavefunctions complete. Namely, we can reverse the projection process to directly determine the bulk wavefunctions from far-field radiations if \(\hat{P}\) is a non-singular invertible 2 × 2 matrix. In this work, we focus on the simplest N = 2 case to present and validate our theory.
Non trivial Berry curvature in honeycomb lattice
To elaborate the correspondence, we consider a 2D PhC slab of Si3N4 material with circular air hole patterns on a honeycomb lattice (Fig. 2A). The radii of two air holes in rhombus unit-cell are denoted as r1 and r2, and the lattice constant and slab thickness are denoted as a and h, respectively. The reciprocal lattice is shown in Fig. 2B, where the gray shading area denotes the reduced BZ. According to the Bloch theorem, the wavefunction \(| {u}_{n}\left.\right\rangle\) can be expanded as a superposition of a series of quasi-plane waves with discrete momentum, represented by the dots in the reciprocal lattice, and we refer to them as “diffraction orders”67. Around the second \({{{\mathcal{K}}}}\) point that resides in the continuum, several diffraction orders fall into the light cone and thus open radiation channels. We take \({{{{\mathcal{K}}}}}_{1}\) point as a specific example, around which exist three radiation channels C1−3 (red arrows, Fig. 2B) whose in-plane wavevectors are \({{{{\boldsymbol{\beta }}}}}_{1}=({k}_{x}+\sqrt{3}/3){\beta }_{0}{{{\bf{x}}}}\) and \({{{{\boldsymbol{\beta }}}}}_{2,3}=({k}_{x}-\sqrt{3}/6){\beta }_{0}{{{\bf{x}}}}+({k}_{y}\;\mp \; 1/2){\beta }_{0}{{{\bf{y}}}}\), respectively. Here, kx and ky are dimensionless numbers describing the Bloch wavevector k∥ deviating from the \({{{{\mathcal{K}}}}}_{1}\) point, and \({\beta }_{0}=4\pi /\sqrt{3}a\) is the reciprocal lattice constant. Three quasi-transverse-electric (TE) polarized modes marked as TEA,B,C with dominated components of (Ex, Ey, Hz) can be found around the \({{{{\mathcal{K}}}}}_{1}\) point due to C6/3 symmetry, forming a three-level system with well-separated frequency from the other Bloch modes. For the case that δr = r2 − r1 = 0 preserving the C6 symmetry (left panel, Fig. 2C), a twofold degeneracy of TEA and TEB right at the \({{{{\mathcal{K}}}}}_{1}\) point emerges. As a contrary, r1 ≠ r2 lifts the degeneracy (right panel, Fig. 2D). In the case that the in-plane symmetry breaking is sufficiently weak that allows the coupling from the TEC mode being neglectable, TEA,B modes form a two-level subsystem near the \({{{{\mathcal{K}}}}}_{1}\) point, described by an effective Hamiltonian of:
in which ω is the degenerate frequency at \({{{{\mathcal{K}}}}}_{1}\) point; δ ∝ δr characterizes the in-plane asymmetry and η is the group velocity. Due to the radiation loss, ω = ωr + iγ0 is a complex one, and γ0 denotes the radiation decay rate. When C6 symmetry is preserved (δ = 0), the eigenvectors of Eq. (2) can be derived as \(| {u}_{n}\left.\right\rangle={[1,\pm | \eta | {e}^{i\theta }/\eta ]}^{T}\) where \({e}^{i\theta }=({k}_{x}+i{k}_{y})/\sqrt{{k}_{x}^{2}+{k}_{y}^{2}}\), creating a diabolic point at the \({{{{\mathcal{K}}}}}_{1}\) point which is exactly a non-Hermitian counterpart of the Dirac point in Hermitian case, and we still denote it as the DP for short without any confusion.
A SEM image of the fabricated PhC sample, showing a honeycomb-latticed structure of SiN slab on SiO2 substrate with different hole radii. Inset: the side view of the air hole. The structural parameters are given as a = 440 nm, r1 = 50 nm, r2 = 54 nm, h = 180 nm, respectively. δr is defined as r2 − r1. B The reciprocal lattice of the PhC sample. Gray shading area: the reduced BZ; purple dot: the Γ point; blue dot: the \({{{{\mathcal{K}}}}}_{1}\) point; red vectors: three diffraction orders acting as radiation channels C1−3; green vectors: non-radiative diffraction orders. C, D The band structures of PhCs around \({{{{\mathcal{K}}}}}_{1}\) point with (δr = 0) and without (δr = 4 nm) the inversion symmetry. Owing to the C6 symmetry with δr = 0, TEA mode (purple sheet) and TEB mode (blue sheet) are degenerate at \({{{{\mathcal{K}}}}}_{1}\) point, giving rise to a DP. When δr ≠ 0, the DP splits to open a nontrivial bandgap. E, F The polarization fields in momentum space of TEA (purple, the top panels) and TEB (blue, the bottom panels) modes around \({{{{\mathcal{K}}}}}_{1}\) point with (δr = 0) and without (δr = 4 nm) the inversion symmetry, respectively. For δr = 0 with the C6 symmetry, half charges emerge at \({{{{\mathcal{K}}}}}_{1}\) point due to the DP for both TEA,B modes. When δr ≠ 0, two CPs with opposite handness emerge around the \({{{{\mathcal{K}}}}}_{1}\) point instead. Black dot: DP; red marks: the quasi-CPs. All data are calculated by numerical simulations (COMSOL Multiphysics).
Degeneracy at DP creates nontrivial bulk band topology. According to Eq. (2), the close form of the theoretical bulk Berry curvature in such a two-level subsystem follows:
where the subscript “t” distinguishes Bn;t from the Bn while the latter one takes TEC into account; the signs “±” denote the two bands n = A, B, respectively. Since TEC mode can be neglected in our case, we have Bn ≈ Bn;t. For a C6 symmetric system (δ = 0), the bulk Berry curvature shows as a δ-function peaked at the \({{{{\mathcal{K}}}}}_{1}\) point (kx = ky = 0). While for δ ≠ 0, the DP splits and opens a topologically nontrivial bandgap, leading to a nontrivial bulk Berry curvature in the vicinity of \({{{{\mathcal{K}}}}}_{1}\) point that gives rise to nonzero valley-Chern number.
As we stated above, if the projection matrix \(\hat{P}\) doesn’t cause extra amplitude vortexes, Bn;t can be determined by directly observing the radiation Berry curvature \({B}_{n;t}^{r}\) according to bulk-radiation correspondence in Eq. (1). To achieve this, both right and left radiation vectors should be measured. Specifically, the right radiation vector \(| {\Psi }_{n}\left.\right\rangle\) corresponds to the far-field polarization: \({\overrightarrow{S}}_{n}={[{s}_{1},{s}_{2},{s}_{3}]}^{T}/{s}_{0}=\langle {\Psi }_{n}| \hat{{{{\mathbf{\sigma }}}}}| {\Psi }_{n}\rangle /\langle {\Psi }_{n}| {\Psi }_{n}\rangle\), where \(\hat{{{{\mathbf{\sigma }}}}}={[{\hat{\sigma }}_{z},{\hat{\sigma }}_{x},{\hat{\sigma }}_{y}]}^{T}\) are the Pauli’s matrices, and \({\overrightarrow{S}}_{n}\) refers to the Stokes’ vector which can be directly measured by using standard polarimetry method65. Here we present the simulated polarization vector fields in Fig. 2E, F, showing the footprints of DP in the far-field radiation with the C6 symmetry. Polarization vortexes can be found in momentum space, each carrying a half-integer topological charge right at the \({{{{\mathcal{K}}}}}_{1}\) point (Fig. 2E). Once the symmetry-breaking lifts the degeneracy, the polarization vortexes degrade to a meron-like and anti-meron-like configurations68 with circular-polarization (CPs) in opposite helicities around the \({{{{\mathcal{K}}}}}_{1}\) point for TEA,B modes, respectively (Fig. 2F). These far-field geometric features can be employed as observable signatures to validate the theory. As for the left radiation vector \(\left\langle \right.{\Phi }_{n}|\), it can be readily determined according to the bi-orthogonal normalization relation of 〈Φm∣Ψn〉 = δmn.
Experimental observation of Berry curvature
To experimentally observe the Berry curvature, we first fabricate the PhC sample by using e-beam lithography (EBL) and inductively coupled plasma etching (ICP) processes on a Si3N4 slab of thickness h = 180 nm on silica substrate (see Methods section for details). The air holes are arranged as a honeycomb lattice of a = 440 nm with two slightly different hole radii of r1 = 50 and r2 = 54 nm, respectively, as the scanning electron microscope (SEM) images shown in Fig. 2A. The angle-resolved measurement system is schematically illustrated in Fig. 3A, in which a supercontinuum white light source is first sent through an acoustic-optic tunable filter (AOTF) and then linearly polarized by POL1 to generate incoherent light in a wavelength range from 550 to 580 nm. After passing through a quarter-wave plate (QWP1), the light is focused by a lens (L1) onto the rear focal plane (RFP) of an infinity-corrected objective lens (NA = 0.95), and then illuminates the sample to excite the optical modes. The radiations from the PhC sample are collected by the same objective lens and imaged by a charge-coupled device (CCD) camera which is co-focused with the RFP of the objective lens. By inserting another polarizer (POL2) and another quarter-wave plate (QWP2) before the CCD, we can fully characterize the Stokes’ vector of radiation through a polarimetry method (see Methods section for details).
A Schematic of the measurements setup. AOTF acoustic-optic tunable filter; POL1 and POL2 polarizers, QWP1 and QWP2 quarter-wave plates, L1 convex lens with 20-cm focal length, BS beam splitter, RFP rear focal plane, Obj objective lens with NA of 0.95 and working distance of 150 μm, 4f lens system with magnification rate of × 6. B The observed image of three scattered beams from radiation channels C1−3 within the NA range. In the experiment, we excite the optical modes from channel C1 by moving lens L1 to a proper position and then collect the radiative waves in channel C3 after it is magnified by the 4f system. C Schematic of isofrequency contours of the two-level system near the \({{{{\mathcal{K}}}}}_{1}\) point. The yellow plane denotes the wavelength of λi = 570.328 nm. Purple sheet: TEA mode; blue sheet: TEB mode. D The measured isofrequency contour S0 and Stokes' parameters S1−3 at a wavelength of λi = 570.328 nm. The Stokes' parameters are determined through different configurations of POL2 and QWP2. Dashed line: the simulated isofrequency contour at λi. E The measured polarization distributions in momentum space around the \({{{{\mathcal{K}}}}}_{1}\) point, by overlapping several isofrequency contours and evaluating the overall Stokes' parameters. Red marks: the quasi-CPs.
As shown in Fig. 3B, three scattered beams inside the aperture of the objective lens can be found, corresponding to the radiation channels C1−3 plotted in Fig. 2B. As we discussed above, only one radiation channel is sufficient to capture the bulk band topology of the two-level subsystem consisting of TEA,B. Therefore, we excite the modes from the C1 channel and observe them from the C3 channel to avoid the direct reflected light for a better signal-noise ratio. The C2 channel is unused. To achieve the best excitation, we fine-tune the incident angle by moving the L1 lens in the x– y plane. Moreover, POL1 and QWP1 are employed to adjust the polarization of incident light: left-handed circular polarization (LCP) for TEA mode and right-handed circular polarization (RCP) for TEB mode, respectively. When the AOTF selects a specific wavelength, the scatterings originating from fabrication disorders would make the isofrequency contour at such a wavelength visible due to the on-resonance pumping mechanism48,62,69 (Fig. 3C). Then, a cascaded 4f system is applied to zoom-in the contour at a magnification rate of × 6. Further, by recording the CCD images at particular arrangements of POL2 and QWP2 as polarimetry measurements, we can obtain the Stokes’ coefficients for each contour. The result at the wavelength of λi = 570.328 nm is presented in Fig. 3D, in which the dashed line is the isofrequency contour calculated from numerical simulation for visual guidance. By overlapping all the isofrequency contours in the wavelength range of 559.383 – 564.861 nm for the TEA band and 565.203 –571.352 nm for the TEB band, we obtain the polarization vector fields for both TEA,B bands (Fig. 3E). In the vicinity of the \({{{{\mathcal{K}}}}}_{1}\) point, an LCP and an RCP can be found on the TEA and TEB bands, respectively (red circles, Fig. 3E), which agree well with our prediction in simulation as shown in Fig. 2F.
The radiation Berry curvature \({B}_{n;t}^{r}\) can be directly obtained from the polarization vector fields measured from channel C3. Here we consider four samples with radius differences of δr = 0, 4, 7, 10 nm, and measure the \({B}_{A,B;t}^{r}\) of each sample (top panels, Fig. 4). As a comparison, we numerically calculate the bulk Berry curvature BA,B by employing the semi-analytical coupled-wave theory framework56,70,71 (bottom panels, Fig. 4). To better show the evolution of Berry curvatures, we plot the unit-cell geometry of each sample and the corresponding band structures of TEA,B as the insets in the top and bottom panels of Fig. 4. Specifically, we start from a realistic sample with δr = 0 (Fig. 4A), where the fabrication imperfections would inevitably break the DP degeneracy at \({{{{\mathcal{K}}}}}_{1}\) point and give rise to a very small bandgap. We estimate that the bandgap is equivalent to the case of δr = 2 nm. In this case, we found that the \({B}_{A,B;t}^{r}\) show as bright spots centered at \({{{{\mathcal{K}}}}}_{1}\) point—quite like δ-functions (top panels, Fig. 4A), agreeing well with the numerical bulk Berry curvatures BA,B (bottom panels, Fig. 4A). Note that the signs of \({B}_{A;t}^{r}\) and \({B}_{B;t}^{r}\) are exactly opposite, agreeing well with the theoretical prediction in Eq. (3). Further, we gradually open the bandgap by increasing δr from 4 to 10 nm (Fig. 4B). During this process, the Berry curvatures \({B}_{B;t}^{r}\) and BB both gradually diffuse to a larger region in momentum space while their peak absolute values decrease. At δr = 10 nm, the bandgap becomes quite large, and thus both \({B}_{n;t}^{r}\) and Bn become fully dispersed and no longer congregate around \({{{{\mathcal{K}}}}}_{1}\) point. \({B}_{A;t}^{r}\) and BA also match well with each other and have similar behaviors (see Methods section). The great agreement between the experimentally observed radiation Berry curvatures \({B}_{A,B;t}^{r}\) and the numerically calculated bulk Berry curvature BA,B confirms the bulk-radiation correspondence we propose in Eq. (2). It’s noteworthy that Berry curvatures are generally complex values in a non-Hermitian system. For our PhC slab in which only radiation contributes to the non-Hermiticity, the imaginary parts of Berry curvature are quite small compared to their real parts (see Supplementary Materials section 1 for details).
A The measured radiation Berry curvatures \({B}_{n;t}^{r}\) for the PhC sample with δr ≈ 0 (top panels) and the numerically calculated bulk Berry curvatures Bn with δr = 2 nm (bottom panels) for a comparison. For a realistic sample with δr = 0, the fabrication errors would slightly lift the DP at \({{{{\mathcal{K}}}}}_{1}\) point to create a small bandgap. We estimate that the bandgap is equivalent to the case of δr = 2 nm. In this case, the Berry curvatures congregate around the \({{{{\mathcal{K}}}}}_{1}\) point since the bandgap is very small, showing opposite signs for TEA and TEB modes. B The measured radiation Berry curvatures \({B}_{B;t}^{r}\) (top panels) and the numerically calculated bulk Berry curvatures BB (bottom panels) for TEB mode when δr = 4 nm (left), 7 nm (middle), and 10 nm (right). Along with the increasing δr, the bandgap gradually opens, and the Berry curvature gradually diffuses to a larger region in momentum space. Insets in top panels: SEM images of the unit cell of each PhC sample; insets in bottom panels: band structures of the two-level system accordingly.
To further quantitatively validate the bulk-radiation correspondence, we calculate the geometric phases (Berry phases) \({\gamma }_{n;t}^{r}\) and γn by applying 2D integrals over the radiation (measured) and bulk (numerical) Berry curvatures of \({B}_{n;t}^{r}\) and Bn in Fig. 4, respectively. As a reference, we also derive the analytic expression of geometric phase γn;t according to Eq. (3), in which the contribution of TEC band is neglected:
According to valleytronics11, the integral of Berry curvature upon the individual valley region determines the valley-Chern number (blue shading, Fig. 5A). Considering that the nontrivial Berry curvatures congregate around the \({{{{\mathcal{K}}}}}_{1}\) point, here we choose a circular integral region with radius of ks = 0.03 for simplicity and thus calculate \({\gamma }_{n;t}^{r}\) (triangle), γn (circle), and γn;t (solid line) shown in Fig. 5B, C. According to Eq.(4), the analytic geometric phases γn;t exactly equals to ±π at δr = 0 for TEA,B bands72, corresponding to the nontrivial quantized valley-Chern number of \({C}_{v}^{{{{\mathcal{K}}}}}=\pm 1/2\) in an individual valley11. When δr ≠ 0, the open bandgap (nonzero δ) would make the geometric phases deviate from ±π, unless the integral region ks tends to be infinite2. Our experimental observation verifies such a behavior. For instance, at δr = 10 nm, we obtain \({\gamma }_{A,B;t}^{r} \approx \pm 0.7\pi\) deviated from ±π due to the impact of nonzero bandgap. On the other hand, we find the three geometric phases \({\gamma }_{n;t}^{r}\), γn and γn;t agree well with each other when δr is relatively small, quantitatively proving the validity of bulk-radiation correspondence. For a large δr, both analytic geometric phase γn;t solved from the two-level model and the measured \({\gamma }_{n;t}^{r}\) from radiation slightly deviate from the bulk Berry phase γn, due to the exclusion of the third TEC band. Because the TEC mode also contributes to far-field radiation, the deviation upon \({\gamma }_{n;t}^{r}\) is more notable than that upon γn;t. Actually, we found \({\gamma }_{B;t}^{r} \, < \,{\gamma }_{B;t}\, \lesssim \,{\gamma }_{B}\) for TEB mode. To improve the measurement accuracy for large δr, TEC mode can’t be neglected, and we need to measure the polarization distributions from two independent radiation channels (i.e., C3 and C2) to completely capture the information of the three-level system with higher DOFs (N = 3). The exemplary example and more discussions are presented in Supplementary Materials section 4 and 5.
A Schematic of an individual valley (blue shading) around the \({{{{\mathcal{K}}}}}_{1}\) point (blue dot) in the reciprocal lattice. The integral on Berry curvature over the individual valley gives the geometric (Berry) phases. Considering that the Berry curvatures congregate around the \({{{{\mathcal{K}}}}}_{1}\) point when δr is relatively small, we perform the integral on a circular region (black circle) to simplify the calculation. Gray shading: the first BZ; purple dot: the Γ point; blue dot: the \({{{{\mathcal{K}}}}}_{1}\) point. B, C The geometric phases for TEA,B modes. Blue solid line: theoretical bulk Berry phase γn;t of the two-level model according to Eq. (4); red circles: numerical bulk Berry phase γn obtained from the integral of bulk Berry curvatures Bn calculated in Fig. 4; yellow triangles: geometric phase \({\gamma }_{n;t}^{r}\) obtained from measured radiation Berry curvature \({B}_{n;t}^{r}\) shown in Fig. 4. When δr = 0, the theoretical Berry phases are exactly ±π for TEA,B modes owing to the existence of DP, corresponding to quantized valley-Chern numbers of ±1/2. When δr gradually increases, all the geometric phases gradually deviate from the quantized ±π. Moreover, when δr becomes relatively large, the theoretical γn;t and measured \({\gamma }_{n;t}^{r}\) slightly deviate from the numerical γn, due to the impact of TEC mode.
Discussion
We emphasize that although the non-Hermiticiy in our exemplary system doesn’t give rise to nontrivial topological effects, the system is still intrinsically non-Hermitian due to the radiation loss, and hence, its mathematical structure must follow a bi-orthogonal relation of left and right vectors together. This indicates that the far-field observables must provide sufficient DOFs to solve the Berry curvatures. Compared to the observables such as wavelength and Q-factor, which are only associated with the complex frequencies of Bloch modes, the far-field polarization is a direct projection of the wavefunction itself that naturally presents the information about it (see Supplementary Materials section 4 for details). Therefore, we conclude that the polarizations should be a better candidate to measure the Berry curvature. Our theory and method are valid for general non-Hermitian photonic systems, even with multi-internal DOFs, as long as the measurable radiation exists.
In the experiment stated above, we directly measure the two right vectors \(| {\Psi }_{A,B}\left.\right\rangle\) at the k∥ point and obtain the left vectors \(\left\langle \right.{\Phi }_{A,B}|\) by using the bi-orthogonal normalization relation. However, the left vector can also be directly measured. According to the law of reciprocity, the left radiation vector at the k∥ point would correspond to the right radiation vector at the − k∥ point (see Supplementary Materials section 3 for details). Therefore, one can also observe one band at both k∥ and − k∥ points to retrieve the system’s topology, instead of measuring two bands simultaneously.
Our findings reveal that the radiation polarization does contain the fingerprint of bulk topology described by the Berry curvature upon a bi-orthogonal basis, but it remains elusive whether all the topological features in the far-field originate from the band topology. Our short-cut answer is negative. Specifically, different from the Berry curvature depending on bi-orthogonal basis, we can also define a “right-right" curvature of \({B}_{n;rr}^{r}=i{\nabla }_{{{{{\bf{k}}}}}_{\parallel }}\times \langle {\Psi }_{n}({{{{\bf{k}}}}}_{\parallel })| {\nabla }_{{{{{\bf{k}}}}}_{\parallel }}{\Psi }_{n}({{{{\bf{k}}}}}_{\parallel })\rangle\) upon the right radiation vector \(| {\Psi }_{n}\left.\right\rangle\) only40, capturing the geometric features of the polarization field itself. In fact, the integral over “right-right" curvature presents the Pancharatnam-Berry (PB) phase73,74,75,76 of far-field polarization, showing the swirling structure of Stokes’ vector \({\overrightarrow{S}}_{n}\) in momentum space. We note that, nontrivial polarization features of meron and anti-meron in Fig. 3E don’t correspond to the valley-Chern number of bulk topology but are related to the Skyrmion numbers77,78,79 given by the PB phases (see Supplementary Materials section 6 for details).
In summary, our findings of “bulk-radiation correspondence” of Berry curvature reveal the feasibility of retrieving the band topology from characterizing far-field radiation. We prove in theory and demonstrate in experiments that employing the polarizations of one radiation channel can accomplish a complete map of wavefunctions in a two-level non-Hermitian system, to directly access the Berry curvature and Chern number without strongly disturbing the system. The proposed method can also be extended to multi-level systems by measuring more radiation channels simultaneously, and utilized to extract other topological features such as quantum geometric tensor25,26,27,28,80,81. Our work demonstrates a simple and effective way of directly observing the Berry curvature in non-Hermitian systems and thus could shed light on the exploration of the intriguing phases in topological systems.
Methods
Sample fabrication
A LPCVD deposited Si3N4-SiO2-Si wafer with 1-μm-thick lower cladding of SiO2 was used. The pattern was defined by electron-beam lithography (EBL). The sample was firstly spin-coated with ZEP520A photo-resist, followed by exposure to EBL (Elionix ELS-F125G8) at the current of 1 nA and field size of 500 μm. After the exposure, the sample was etched with inductively coupled plasma (ICP, Oxford) by a mixture of CHF3 and O2. The ICP etching time and chamber pressure are carefully controlled for etching depth and smooth side walls.
Measurements
As shown in Fig. 6, the supercontinuum light source (SuperK Compact; NKT Photonics) is broadband in the range of 450–2400 nm. The incident light is followed by an acoustic-optic tunable filter (AOTF; Gooch & Housego) to scan and pick up the resonance wavelength. After passing through a polarizer (POL1) and a quarter-wave plate (QWP1), the incident light is focused by a lens (L1) onto the rear focal plane (RFP) of the objective lens for exciting the sample. We control the incident light to be left-handed and right-handed circular polarized by arranging POL1 and QWP1 to excite TEA and TEB bands, respectively. The radiation field of the sample, collected by the objective lens, is magnified by a two-stage 4f system (L2– L5) and passes through a quarter-wave plate (QWP2) and a polarizer (POL2) before observed by a CCD camera (CCD1). Another 4f system (L6 – L7) and camera (CCD2) are used to monitor the near field of the sample surface. We perform polarimetry measurement of the diffraction from the C3 radiation channel by six configurations of QWP2 and POL2: (1) no QWP2, POL2 oriented along the x-axis; (2) no QWP2, POL2 oriented along the y-axis; (3) no QWP2, POL2 oriented at 45° with respect to the x-axis; (4) QWP2 fast axis fixed along y-axis, POL2 oriented at 45° with respect to the x-axis; (5) no QWP2, POL2 oriented at 135° with respect to the x-axis; (6) QWP2 fast axis fixed along y-axis, POL2 oriented at 135° with respect to the x-axis. The far-field images are recorded by a CCD camera for further processing.
Measurement results of the Stokes’ parameters
We present the measurement results of the Stokes’ parameters of both TEA and TEB bands for the PhC slab of δr = 4 nm, shown in Fig. 7. The experiment and the simulation results match well with each other.
Left panels: Measured (top panels) and simulated (bottom panels) Stokes' parameters (S1−3) for TEA mode when δr = 4; Right panels: Measured (top panels) and simulated (bottom panels) Stokes' parameters (S1−3) for TEB mode when δr = 4. For both two modes, the experiment results agree well with the simulated ones.
Berry curvature for TEA band
We present the evolution of measured radiation Berry curvature for TEA band when δr varies in Fig. 8. Also, the experiment results agree well with the simulation results (bulk Berry curvature).
Top panels: Measured radiation Berry curvature \({B}_{A;t}^{r}\) when δr ≈ 0 (the first from left), δr = 4 nm (the second from left), 7 nm (the third from left), and 10 nm (right). Bottom panels: simulated bulk Berry curvature BA when δr = 2 nm (the first from left), δr = 4 nm (the second from left), 7 nm (the third from left), and 10 nm (right). The experimental results agree well with the simulated ones. Along with the increasing δr, the bandgap between TEA mode and TEB mode increases, and the Berry curvatures gradually diffuse to a larger region in momentum space.
Data availability
All data needed to evaluate the conclusions in the paper are present in the paper and the supplementary materials.
References
Hasan, M. Z. & Kane, C. L. Colloquium: topological insulators. Rev. Mod. Phys. 82, 3045 (2010).
Xiao, D., Chang, M.-C. & Niu, Q. Berry phase effects on electronic properties. Rev. Mod. Phys. 82, 1959 (2010).
Lu, L., Joannopoulos, J. D. & Soljačić, M. Topological photonics. Nat. Photon. 8, 821–829 (2014).
Khanikaev, A. B. & Shvets, G. Two-dimensional topological photonics. Nat. Photon. 11, 763–773 (2017).
Ozawa, T. et al. Topological photonics. Rev. Mod. Phys. 91, 015006 (2019).
Bergholtz, E. J., Budich, J. C. & Kunst, F. K. Exceptional topology of non-hermitian systems. Rev. Mod. Phys. 93, 015005 (2021).
Berry, M. V. Quantal phase factors accompanying adiabatic changes. Proc. R. Soc. A 392, 45–57 (1984).
Haldane, F. D. M. Berry curvature on the fermi surface: anomalous hall effect as a topological fermi-liquid property. Phys. Rev. Lett. 93, 206602 (2004).
Thouless, D. J., Kohmoto, M., Nightingale, M. P. & den Nijs, M. Quantized hall conductance in a two-dimensional periodic potential. Phys. Rev. Lett. 49, 405 (1982).
Sheng, D., Weng, Z., Sheng, L. & Haldane, F. Quantum spin-hall effect and topologically invariant chern numbers. Phys. Rev. Lett. 97, 036808 (2006).
Xiao, D., Yao, W. & Niu, Q. Valley-contrasting physics in graphene: magnetic moment and topological transport. Phys. Rev. Lett. 99, 236809 (2007).
Zhang, F., Jung, J., Fiete, G. A., Niu, Q. & MacDonald, A. H. Spontaneous quantum hall states in chirally stacked few-layer graphene systems. Phys. Rev. Lett. 106, 156801 (2011).
Zhang, F., MacDonald, A. H. & Mele, E. J. Valley chern numbers and boundary modes in gapped bilayer graphene. Proc. Natl Acad. Sci. USA 110, 10546–10551 (2013).
Hauke, P., Lewenstein, M. & Eckardt, A. Tomography of band insulators from quench dynamics. Phys. Rev. Lett. 113, 045303 (2014).
Fläschner, N. et al. Experimental reconstruction of the berry curvature in a floquet bloch band. Science 352, 1091–1094 (2016).
Price, H. M. & Cooper, N. Mapping the berry curvature from semiclassical dynamics in optical lattices. Phys. Rev. A 85, 033620 (2012).
Jotzu, G. et al. Experimental realization of the topological haldane model with ultracold fermions. Nature 515, 237–240 (2014).
Aidelsburger, M. et al. Measuring the chern number of hofstadter bands with ultracold bosonic atoms. Nat. Phys. 11, 162–166 (2015).
Ozawa, T. & Carusotto, I. Anomalous and quantum hall effects in lossy photonic lattices. Phys. Rev. Lett. 112, 133902 (2014).
Wimmer, M., Price, H. M., Carusotto, I. & Peschel, U. Experimental measurement of the berry curvature from anomalous transport. Nat. Phys. 13, 545–550 (2017).
Abanin, D. A., Kitagawa, T., Bloch, I. & Demler, E. Interferometric approach to measuring band topology in 2d optical lattices. Phys. Rev. Lett. 110, 165304 (2013).
Atala, M. et al. Direct measurement of the zak phase in topological bloch bands. Nat. Phys. 9, 795–800 (2013).
Duca, L. et al. An Aharonov-Bohm interferometer for determining bloch band topology. Science 347, 288–292 (2015).
Li, T. et al. Bloch state tomography using Wilson lines. Science 352, 1094–1097 (2016).
Bleu, O., Solnyshkov, D. & Malpuech, G. Measuring the quantum geometric tensor in two-dimensional photonic and exciton-polariton systems. Phys. Rev. B 97, 195422 (2018).
Gianfrate, A. et al. Measurement of the quantum geometric tensor and of the anomalous hall drift. Nature 578, 381–385 (2020).
Ren, J. et al. Nontrivial band geometry in an optically active system. Nat. Commun. 12, 689 (2021).
Liao, Q. et al. Experimental measurement of the divergent quantum metric of an exceptional point. Phys. Rev. Lett. 127, 107402 (2021).
Polimeno, L. et al. Tuning of the berry curvature in 2d perovskite polaritons. Nat. Nanotechnol. 16, 1349–1354 (2021).
Łempicka-Mirek, K. et al. Electrically tunable berry curvature and strong light-matter coupling in liquid crystal microcavities with 2d perovskite. Sci. Adv. 8, eabq7533 (2022).
Wu, S. et al. Electrical tuning of valley magnetic moment through symmetry control in bilayer mos2. Nat. Phys. 9, 149–153 (2013).
Cho, S. et al. Experimental observation of hidden berry curvature in inversion-symmetric bulk 2 h- wse 2. Phys. Rev. Lett. 121, 186401 (2018).
Vallone, G. & Dequal, D. Strong measurements give a better direct measurement of the quantum wave function. Phys. Rev. Lett. 116, 040502 (2016).
Dressel, J., Malik, M., Miatto, F. M., Jordan, A. N. & Boyd, R. W. Colloquium: understanding quantum weak values: basics and applications. Rev. Mod. Phys. 86, 307 (2014).
Leykam, D. & Smirnova, D. A. Probing bulk topological invariants using leaky photonic lattices. Nat. Phys. 17, 632–638 (2021).
Gorlach, M. A. et al. Far-field probing of leaky topological states in all-dielectric metasurfaces. Nat. Commun. 9, 909 (2018).
Feng, L., El-Ganainy, R. & Ge, L. Non-hermitian photonics based on parity–time symmetry. Nat. Photon. 11, 752–762 (2017).
Leykam, D., Bliokh, K. Y., Huang, C., Chong, Y. D. & Nori, F. Edge modes, degeneracies, and topological numbers in non-hermitian systems. Phys. Rev. Lett. 118, 040401 (2017).
El-Ganainy, R. et al. Non-Hermitian physics and PT symmetry. Nat. Phys. 14, 11–19 (2018).
Shen, H., Zhen, B. & Fu, L. Topological band theory for non-Hermitian Hamiltonians. Phys. Rev. Lett. 120, 146402 (2018).
Zhen, B., Hsu, C. W., Lu, L., Stone, A. D. & Soljačić, M. Topological nature of optical bound states in the continuum. Phys. Rev. Lett. 113, 257401 (2014).
Doeleman, H. M., Monticone, F., den Hollander, W., Alu, A. & Koenderink, A. F. Experimental observation of a polarization vortex at an optical bound state in the continuum. Nat. Photon. 12, 397–401 (2018).
Zhang, Y. et al. Observation of polarization vortices in momentum space. Phys. Rev. Lett. 120, 186103 (2018).
Chen, W., Chen, Y. & Liu, W. Singularities and poincaré indices of electromagnetic multipoles. Phys. Rev. Lett. 122, 153907 (2019).
Yin, X. & Peng, C. Manipulating light radiation from a topological perspective. Photonics Res. 8, B25–B38 (2020).
Liu, W., Liu, W., Shi, L. & Kivshar, Y. Topological polarization singularities in metaphotonics. Nanophotonics 10, 1469–1486 (2021).
Che, Z. et al. Polarization singularities of photonic quasicrystals in momentum space. Phys. Rev. Lett. 127, 043901 (2021).
Zhou, H. et al. Observation of bulk fermi arc and polarization half charge from paired exceptional points. Science 359, 1009–1012 (2018).
Chen, W., Yang, Q., Chen, Y. & Liu, W. Evolution and global charge conservation for polarization singularities emerging from non-hermitian degeneracies. Proc. Natl Acad. Sci. USA 118, e2019578118 (2021).
Huang, C. et al. Ultrafast control of vortex microlasers. Science 367, 1018–1021 (2020).
Zhang, X., Liu, Y., Han, J., Kivshar, Y. & Song, Q. Chiral emission from resonant metasurfaces. Science 377, 1215–1218 (2022).
Chen, Y. et al. Observation of intrinsic chiral bound states in the continuum. Nature 613, 474–478 (2023).
von Neuman, J. & Wigner, E. Über merkwürdige diskrete Eigenwerte. Uber das Verhalten von Eigenwerten bei adiabatischen Prozessen. Physikalische Z. 30, 467–470 (1929).
Friedrich, H. & Wintgen, D. Interfering resonances and bound states in the continuum. Phys. Rev. A 32, 3231–3242 (1985).
Hsu, C. W. et al. Observation of trapped light within the radiation continuum. Nature 499, 188–191 (2013).
Yang, Y., Peng, C., Liang, Y., Li, Z. & Noda, S. Analytical perspective for bound states in the continuum in photonic crystal slabs. Phys. Rev. Lett. 113, 037401 (2014).
Hsu, C. W., Zhen, B., Stone, A. D., Joannopoulos, J. D. & Soljačić, M. Bound states in the continuum. Nat. Rev. Mater. 1, 16048 (2016).
Jin, J. et al. Topologically enabled ultrahigh-Q guided resonances robust to out-of-plane scattering. Nature 574, 501–504 (2019).
Sadreev, A. F. Interference traps waves in an open system: bound states in the continuum. Rep. Prog. Phys. 84, 055901 (2021).
Kang, M., Zhang, S., Xiao, M. & Xu, H. Merging bound states in the continuum at off-high symmetry points. Phys. Rev. Lett. 126, 117402 (2021).
Hu, P. et al. Global phase diagram of bound states in the continuum. Optica 9, 1353–1361 (2022).
Yin, X., Jin, J., Soljačić, M., Peng, C. & Zhen, B. Observation of topologically enabled unidirectional guided resonances. Nature 580, 467–471 (2020).
Zeng, Y., Hu, G., Liu, K., Tang, Z. & Qiu, C.-W. Dynamics of topological polarization singularity in momentum space. Phys. Rev. Lett. 127, 176101 (2021).
Yin, X., Inoue, T., Peng, C. & Noda, S. Topological unidirectional guided resonances emerged from interband coupling. Phys. Rev. Lett. 130, 056401 (2023).
Mcmaster, W. H. Polarization and the Stokes parameters. Am. J. Phys. 22, 351–362 (1954).
Simon, B. Holonomy, the quantum adiabatic theorem, and berry’s phase. Phys. Rev. Lett. 51, 2167 (1983).
Kogelnik, H. & Shank, C. V. Coupled-wave theory of distributed feedback lasers. J. Appl. Phys. 43, 2327–2335 (1972).
Guo, C., Xiao, M., Guo, Y., Yuan, L. & Fan, S. Meron spin textures in momentum space. Phys. Rev. Lett. 124, 106103 (2020).
Regan, E. C. et al. Direct imaging of isofrequency contours in photonic structures. Sci. Adv. 2, e1601591 (2016).
Liang, Y., Peng, C., Sakai, K., Iwahashi, S. & Noda, S. Three-dimensional coupled-wave model for square-lattice photonic crystal lasers with transverse electric polarization: a general approach. Phys. Rev. B 84, 195119 (2011).
Peng, C., Liang, Y., Sakai, K., Iwahashi, S. & Noda, S. Three-dimensional coupled-wave theory analysis of a centered-rectangular lattice photonic crystal laser with a transverse-electric-like mode. Phys. Rev. B 86, 035108 (2012).
Zhang, Y., Tan, Y.-W., Stormer, H. L. & Kim, P. Experimental observation of the quantum hall effect and berry’s phase in graphene. Nature 438, 201–204 (2005).
Pancharatnam, S. Generalized theory of interference, and its applications: part I. coherent pencils. Proc. Indian Acad. Sci. A 44, 247–262 (1956).
Berry, M. V. The adiabatic phase and pancharatnam’s phase for polarized light. J. Mod. Opt. 34, 1401–1407 (1987).
Lee, Y.-H. et al. Recent progress in Pancharatnam–Berry phase optical elements and the applications for virtual/augmented realities. Opt. Data Process. Storage 3, 79–88 (2017).
Xie, X. et al. Generalized Pancharatnam-Berry phase in rotationally symmetric meta-atoms. Phys. Rev. Lett. 126, 183902 (2021).
Skyrme, T. Particle states of a quantized meson field. Proc. Math. Phys. Eng. 262, 237–245 (1961).
Skyrme, T. H. R. A unified field theory of mesons and baryons. Nucl. Phys. 31, 556–569 (1962).
Fert, A., Reyren, N. & Cros, V. Magnetic skyrmions: advances in physics and potential applications. Nat. Rev. Mat. 2, 1–15 (2017).
Provost, J. & Vallee, G. Riemannian structure on manifolds of quantum states. Commun. Math. Phys. 76, 289–301 (1980).
Anandan, J. & Aharonov, Y. Geometry of quantum evolution. Phys. Rev. Lett. 65, 1697 (1990).
Acknowledgements
We acknowledge stimulating discussions with Dr. Takuya Inoue, Prof. Bo Zhen, and Dr. Jicheng Jin. This work was supported from National Key Research and Development Program of China (2022YFA1404804), National Natural Science Foundation of China (62325501 and 62135001), Grant-in-Aid for Scientific Research (22H04915).
Author information
Authors and Affiliations
Contributions
X.Y. and Y.C. contributed equally to this work. X.Y. and C.P. wrote the manuscript with contributions from all authors. X.Y., C.P., and S.N. supervised the project. X.Y. designed the samples and implemented the simulations. Y.C. and Z.Z. conducted the experiments. X.Z. fabricated the samples. X.Y. and Y.C. dealt with the experiment results.
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
Publisher’s 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 article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s 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
Yin, X., Chen, Y., Zhang, X. et al. Observation of Berry curvature in non-Hermitian system from far-field radiation. Nat Commun 16, 2796 (2025). https://doi.org/10.1038/s41467-025-58050-8
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-58050-8