Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Non-Hermitian topological photonics

Open Access Open Access

Abstract

Recent years have witnessed a flurry of research activities in topological photonics, predominantly driven by the prospect for topological protection–a property that endows such systems with robustness against local defects, disorder, and perturbations. This field emerged in fermionic environments and primarily evolved within the framework of quantum mechanics which is by nature a Hermitian theory. However, in light of the ubiquitous presence of non-Hermiticity in a host of natural and artificial settings, one of the most pressing questions today is how non-Hermiticity may affect some of the predominant features of topological arrangements and whether or not novel topological phases may arise in non-conservative and out of equilibrium systems that are open to the environment. Here, we provide a brief overview of recent developments and ongoing efforts in this field and present our perspective on future directions and potential challenges. Special attention will be given to the interplay of topology and non-Hermiticity–an aspect that could open up new frontiers in physical sciences and could lead to promising opportunities in terms of applications in various disciplines of photonics.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Decades of progress carried out independently in non-Hermitian physics and topological insulators have met each other in the flourishing field of non-Hermitian topological photonics [122]. Studies in this area have been stimulated by the observations of counterintuitive topological phenomena that have no counterpart whatsoever in the Hermitian domain. Promising prospects in real-world open systems along with a flexibility for accommodating a variety of non-Hermitian sources on photonic platforms are among the most prominent driving forces in advancing this field. By deploying an arsenal of theoretical, numerical, and experimental tools and methodologies, research in non-Hermitian photonics is nowadays marching through a terrain that extends from mathematics and condensed matter physics to material science, and optics.

Symmetry is one of the essential cornerstones in topology given that it lays the foundations in understanding different topological phases. Classification schemes based on non-spatial internal symmetries are culminated in the celebrated Altland-Zirnbauer systems. These tenfold symmetry classes are categorized in terms of the presence or absence of time-reversal symmetry, particle-hole symmetry, and chiral symmetry. Even though the Altland-Zirnbauer scheme is well-defined in the Hermitian domain and it can be generalized to include spatial symmetries like for example that of reflection, understanding the implications of symmetry in non-Hermitian topological configurations is still a subject of intense investigation [2329]. For instance, it has been found that in the presence of non-Hermiticity, distinct internal symmetries can be unified [30] and sublattice and chiral symmetries no longer coincide [24]. Non-Hermitian systems exhibiting exotic symmetries such as that of parity-time (PT) have no analogue in the Hermitian domain. A non-Hermitian Hamiltonian respects PT symmetry provided that its complex potential V satisfies the condition $V(x )= {V^\ast }({ - x} )$, which is necessary but not sufficient for supporting real eigenvalues. In this regard, one can switch between a PT-symmetric phase, exhibiting real eigenvalues, and a PT-broken phase associated with complex eigenspectra, a phase transition dictated by the magnitude of the complex potential V. To implement such PT ‘optical potentials’ in photonic platforms, an antisymmetric gain/loss profile with symmetric index guiding can be judiciously employed [1,3,3135]. Enhanced modulation bandwidth in PT symmetric coupled lasers reveals how new paradigms in controlling laser characteristics can be introduced by such a non-Hermitian symmetry [36]. Exceptional semimetal–like crystals can be stabilized through novel types of spatial symmetries that are unique to non-Hermitian platforms [37]. These results suggest that there is a drastic departure from what one could expect in conservative systems to what may ensue in the presence of non-Hermiticity [24,25]. On another front, periodically driven non-Hermitian systems can display topological phases that cannot be attained under static conditions. The associated Floquet non-Hermitian topological bands in such structures have been systematically classified according to the internal symmetries based on the so-called K-theory [38].

Different classes of topological non-Hermitian systems have been also extended to include defects [39]. In this context, the existence of topologically protected edge states in PT symmetric arrangements, has been a matter of debate [4043]. Such states were experimentally observed for the first time in passive Su–Schrieffer–Heeger (SSH) lattices inscribed in silica where the waveguide elements were intentionally wiggled as a function of propagation distance in order to fine-tune the loss [44]. Within the framework of time-binding formalism, the following Hamiltonian can be used to describe the dynamics of such SSH configurations:

$$H = {\epsilon _a}\mathop \sum \nolimits_n \hat{c}_n^{a\dagger }\hat{c}_n^a + {\epsilon _b}\mathop \sum \nolimits_n \hat{c}_n^{b\dagger }\hat{c}_n^b + \mathop \sum \nolimits_n [{{\kappa_1}({\hat{c}_n^{b\dagger }\hat{c}_n^a + \hat{c}_n^{a\dagger }\hat{c}_n^b} )+ {\kappa_2}({\hat{c}_{n - 1}^{b\dagger }\hat{c}_n^a + \hat{c}_n^{a\dagger }\hat{c}_{n - 1}^b} )} ]$$
where ${\epsilon _a}$ and ${\epsilon _b}$ denote the complex onsite potentials and $\hat{c}_n^{a\dagger }({\hat{c}_n^a} )$ and $\hat{c}_n^{b\dagger }({\hat{c}_n^b} )$ stand for the associated boson creation (annihilation) operators in the sublattices a and b of this lattice and intracell and intercell hopping strengths are represented by ${\kappa _1}$ and ${\kappa _2}$. Along similar lines, a selective enhancement (or suppression) of robust modes existing at the interface of PT-symmetric microwave SSH lattices has been achieved by introducing loss that breaks the sublattice symmetry [45]. Similarly, the combined effect of non-Hermitian PT and charge-conjugation symmetry in leaky coupled resonator optical waveguides has been shown to produce topologically protected defect states even when their conservative limit is trivial [46]. In another setting, Takata and Notomi proposed a topological phase, solely induced by imaginary potentials, in a lattice composed of four coupled cavities with dimerized gain and loss components [47]. In addition, it has been found that in non-Hermitian Aubry-Andre-Harper (AAH) models, edge states with purely imaginary eigenenergies can arise, that are stabilized by a non-Hermitian particle-hole symmetry– features that are of particular interest to topological mode selection in lasers [48]. The concurrence of triple phase transitions, namely localization, topological and parity–time symmetry breaking, in a temporally driven dissipative synthetic quasicrystal is yet another example of counterintuitive phenomena enabled by the intertwinement of Floquet engineering, topology and non-Hermiticity [49]. Breaking time reversal symmetry in Hermitian topological Chern insulators is typically attained either through gyromagnetic effects [50,51] or temporal modulation [5255]. Yet until recently, asymmetric long-range hopping, first proposed by Duncan Haldane in his seminal paper [17], has evaded experimental observation. Lately, Liu et al. showed how the interplay of non-Hermiticity, and optical gain can be leveraged to tailor the complex next nearest neighbor exchange coupling between judiciously engineered ring resonator elements. The topological response of this system relies on the presence of gain and spontaneous emissions, while it vanishes in its conservative and passive analogue. This active lattice promotes robust edge states and displays single-mode unidirectional topological lasing above the threshold. However, it requires a change of the length of the incorporated links and chirality of the fan-shaped structure in order to induce the associated phase shifts for reversing the flow of light (Fig. 1(a)) [56,57].

 figure: Fig. 1.

Fig. 1. Topological lasers. (a) Robust lasing in an edge mode in a photonic implementation of the Haldane lattice [56]. Lasing in the protected midgap state of a Su–Schrieffer–Heeger (SSH) lattice implemented using (b) a zigzag chain of coupled polariton micropillars [59] and (c) a microring resonators InGaAsP/InP structure [60]. (d) A topological insulator laser using a two-dimensional array of site and intermediary link resonators that emulates the quantum spin Hall effect [64]. (e) Topological lasing in a quantum Hall platform realized by integrating a photonic crystal with a gyromagnetic substrate [68] (f) Topological insulator vertical-cavity surface-emitting laser (VCSEL) array [72]. (g) Multipolar lasing from corner topological modes in a generalized 2D SSH lattice [81]. (h) An electrically pumped topological insulator laser using a quantum spin Hall platform [87]. (i) topological temporal mode-locking between the temporal modes of a fiber lase cavity [88].

Download Full Size | PDF

2. Topological lasers

Such topological lasers, indeed, constitute one of the most technologically relevant settings where topology synergistically interacts with non-Hermiticity and nonlinearity. Lasers, coherent light sources that rely on nonlinear gain saturation, provide a promising test bed for exploring the ramifications of non-Hermiticity and nonlinearity on topology. Along different lines, robust topological lasing was found to mitigate vulnerabilities because of disorder, defects, and operational malfunction – an aspect that makes topological coupled arrays of lasers even more appealing. As such, a vast body of theoretical and experimental studies has been devoted to this topic–reshaping our understanding of non-Hermitian topological systems. Research in this field was initiated by theoretical studies of the SSH model [58]. This was followed by three experiments demonstrating lasing in the associated topologically protected midgap defect state. St.-Jean et al. implemented an orbital version of the SSH Hamiltonian using a zigzag chain of coupled polariton micropillars (Fig. 1(b)) [59]. Parto et al. employed an array of coupled mirroring resonators fabricated on InGaAsP quantum wells that are subject to an optical pumping scheme that respects PT symmetry. Unlike the passive SSH system possessing chiral symmetry, the ensuing Hamiltonian in the non-Hermitian regime has acquired a chiral-time symmetry (Fig. 1(c)) [60]. In a parallel work, Zhao et al. reported lasing of a protected zero mode in coupled InGaAsP-silicon microring resonators on a silicon-on-insulator substrate [61]. Aside from these microscale demonstrations, the SSH model has been realized and lasing from its zero-dimensional topological states has been observed in photonic crystals with embedded nanocavities – a feature of particular interest in miniaturization and high-density integration schemes. In general, the lasing mode can survive onsite disorder as well as perturbations that affect the coupling strength, however, the spectral position of the zero-energy states is only immune to the latter [62,63].

Clearly, of importance would be to identify two–dimensional topological arrangements where robust lasing can be manifested through truly protected edge states. Unlike one-dimensional systems, in 2D configurations this can be achieved via protected unidirectional transport that is akin to that observed in actual fermionic systems. In this regard, Bandres et al. demonstrated robust edge mode lasing along the perimeter of a judiciously designed square lattice of coupled microcavities that employs a synthetic gauge magnetic field for photons. In this vein, it exhibits a quantum spin Hall effect where the unidirectional circulation of photons in the ring resonators (clockwise or counterclockwise) acts as a pseudo-spin. Even at pump powers significantly above threshold, the topological features of this system allow for single mode lasing with a slope efficiency considerably higher than that of the corresponding topologically trivial realization (Fig. 1(d)) [6467]. Around the same time, Bahari et al. built a square photonic crystal on an yttrium iron garnet (YIG) substrate to break time reversal symmetry. The resulting Chern insulator was then interfaced with a topologically trivial triangular crystal to promote unidirectional lasing edge modes under optical pumping (Fig. 1(e)) [68,69]. Along these lines, Klembt et al. investigated lasing in topological edge states in a polariton honeycomb array in the presence of an external magnetic field through the Zeeman effect [70]. Moreover, Dikopoltsev et al. reported a topological insulator vertical-cavity surface-emitting laser (VCSEL) array. In this structure, VCSELs–each of them a circular micropillar–are arranged to form a topological interface between shrunken and expanded honeycomb lattices. The topological features of the array enable the injection locking of VCSELs along the interface–thus making them to lase coherently and operate like a single laser (Fig. 1(f)) [71,72]. In this spirit, proposals were made where a train of mode-locked pulses can be produced by forcing an ensemble of topological semiconductor laser resonators to mutually lock in a synthetic dimension [73]. Alternatively, topological insulator lasers have been realized by exploiting the valley degree of freedom in nanostructured photonic crystals [74]. Interestingly, it has also been suggested that the band-inversion-induced reflection in architected topological nanocavity arrays can be employed as a new type of feedback for developing topological bulk lasers [75]. Gain saturation has been shown to have a dramatic effect on topological insulator lasers by lifting the degeneracy of the clockwise/counterclockwise spin-momentum locked edge states–an aspect enabling single-mode topological vortex lasing in a cross shaped photonic crystal cavity [76].

Recently emerged higher-order topological phases are yet another theme that has generated a great deal of research in the quest for low threshold topological lasers at nanoscale [77,78]. Topological nanocavities based on corner states supported by the generalized 2D SSH model have so far been explored in this pursuit [79,80]. The radiative coupling between the topological corner states results in multipole corner lasing modes with distinct emission patterns [81]. Using an InGaAsP slab with embedded quantum wells as the gain medium [81] instead of quantum dots [79] enables the system to function at room temperatures (Fig. 1(g)) [81]. As an analog to Majorana bound states (MBS) in superconductors, Dirac-vortex states have opened up a new route to two-dimensional topological mid-gap defect cavities. Imposing generalized Kekulé modulations to honeycomb photonic crystals leads to a vortex gap – a photonic realization of Jackiw–Rossi zero modes in 2D Dirac and Hou–Chamon–Mudry (HCM) models used in condensed matter physics. The free spectral range of this Dirac-vortex cavity is found to be larger than any other optical resonator with scalable modal volume – an intriguing attribute for high-power single-mode surface-emitting lasers. The topological protection ensures that the lasing wavelengths are always near the Dirac frequency, regardless of cavity size [8285].

In addition to these optically pumped lasing schemes, electrically pumped topological lasers have been investigated recently in an effort to eliminate the need for an external laser source. Examples includes demonstrations that have been made using topologically protected valley edge modes in the terahertz band [86] and a quantum spin Hall platform in the telecommunication window (Fig. 1(h)) [87]. While the former works at cryogenic temperatures, the latter operated at a room temperature. The robust transport of light along a triangular topological valley cavity without any morphological optimization of the corners and with sharp bends bypassed by the lasing edge mode, is a clear manifestation of topological protection in non-Hermitian systems. In an attempt to introduce topology into mode-locked lasers, Leefmans et al. reported how asymmetric dissipative coupling between temporal modes of a fiber laser cavity can conspire with gain saturation nonlinearity to attain a topological temporal mode-locking. Equidistant temporal modes in such a laser cavity are generated by sinusoidal modulation of the intracavity loss using intensity modulators, while two optical delay lines introduce asymmetric dissipative couplings between nearest-neighbor pulses. Given that the gain medium provided by erbium doped fiber amplifier (EDFA) saturates due to the average power of temporal cavity modes, gain saturation nonlinearity establishes nonlocal interactions between the laser pulses. This technique has been used to demonstrate a nonlinearity driven non-Hermitian skin effect that is fundamentally different from current conventional mechanisms that rely on linear transport processes (Fig. 1(i)) [88].

3. Non-Hermitian bulk-edge correspondence and non-Hermitian skin effect

Localization of all eigenstates at one of the open edges of finite-size (or semi-infinite) and non-Hermitian lattices, with asymmetric couplings as the underlying cause for non-Hermiticity, shows once again how the resulting non-Hermitian skin effect can challenge common wisdom about the topological bulk-edge correspondence [8992]. The counterintuitive features of the non-Hermitian skin effect have sparked two main lines of research over the past few years. The first one focuses on developing non-Bloch frameworks that describe the topology of non-Hermitian bands [93,94]. Biorthogonal bulk-edge correspondence is one such example that replaces conventional eigensystems with a biorthogonal set of right and left eigenstates [95]. Meanwhile, the second one explains the nontrivial topological properties of non-Hermitian arrangements through their complex line and point gap structures–features that are associated with energy bands that can avoid the crossing of a reference line or point in the complex energy plane, respectively. In the presence of a line gap, the Hermitian topological phase has been shown to be resilient to non-Hermitian perturbations, as is the case for a group of topological lasers [64,68,87]. Note that the point gap is unique to non-Hermitian environments and is responsible for the topological origin of the non-Hermitian skin effect [24,25,96,97]. Along similar lines, auxiliary generalized Brillouin zone theory has been established as a self-consistent method to understand bulk-boundary correspondence in one dimensional non-Hermitian configurations [98]. Correspondingly, the topological invariants are either adapted from the Hermitian domain and then generalized to account for the non-Hermiticity or entirely new invariants are developed like for example that of vorticity. In the latter approach, topological invariants are defined not only based on the eigenfunctions but also on the associated complex eigenvalues. On the other hand, the non-Bloch Chern number was found to rely on the non-Bloch nature of eigenstates given that it accurately captures the topological phase transitions in the non-Hermitian regime [24,99]. In this context, the topological charge of the bound states in the continuum (BICs)– as polarization vortices in momentum space –is quantified with a winding number which is not conserved during band conversion. As such, a Skyrmion number has been introduced to describe the topology of symmetry-protected BICs experiencing band inversion in leaky photonic crystals [100].

The non-Hermitian skin effect was introduced in 1990s, when Hatano and Nelson proposed an archetypical disordered tight-binding lattice with asymmetric hoppings – as a non-Hermitian extension of the Anderson model. A mobility edge in this model appears due to the competition between non-Hermiticity and disorder and can be interpreted using the associated point gap topology. Indeed, as a canonical example where topology is induced by non-Hermiticity, the Hatano-Nelson model, possesses a non-trivial (trivial) winding number when the wave functions are extended (localized) [24,101,102]. The first experimental demonstration of the non-Hermitian skin effect was carried out in a time-synthetic photonic mesh lattice composed of two coupled fiber loops (Fig. 2(a)) [103]. In a parallel study, a single-photon interferometer was used that allows discrete-time non-unitary quantum-walk dynamics in order to explore the topological properties of the non-Hermitian skin effect [104]. Subsequently, a Hatano-Nelson model was implemented in a frequency synthetic dimension by simultaneously modulating the resonant modes in a fiber ring cavity in both amplitude and phase [105]. On the other hand, the first-ever realization of the photonic Hatano-Nelson model in real space has been achieved only recently using an array of unidirectional ring resonators and judiciously designed links, where the interplay of optical gain and non-Hermiticity provides a mechanism for asymmetric coupling [106]. The ensued non-Hermitian skin effect in this system has been successfully utilized for phase locking in laser arrays (Fig. 2(b)) [106].

 figure: Fig. 2.

Fig. 2. Non-Hermitian skin effect. (a). Efficient funnel for light based on non-Hermitian skin effect in time synthetic photonic mesh lattices [103]. (b). Unidirectional microring laser with adjustable asymmetric coupling that emulates a Hatano-Nelson model [106].

Download Full Size | PDF

In terms of applications, the sensitivity of the energy spectrum of non-Hermitian Hamiltonians to boundary conditions has been suggested as a means for developing an entirely new class of topological sensors [107109]. This sensitivity originates from the violation of the bulk-boundary correspondence in non-Hermitian environments that is known to grow exponentially with system size [107109]. Parallel to these one-dimensional photonic demonstrations, there have been ongoing efforts in electric circuits [110], acoustics [111,112], mechanics [113], and cold atom systems [114] in producing non-Hermitian skin effects in higher dimensions. In electric circuits, it has been shown that the interaction between multiple and distinct directional hoppings can significantly change the localization of the eigenstates and even lead to complete disappearance of the non-Hermitian skin effect [115]. Of interest, the interplay between the non-Hermitian skin effect and dissipative gaps has unveiled an unexpected boundary-induced dynamical phenomenon. This so called non-Hermitian edge burst is robust against edge perturbations and is manifested as a pronounced peak [116].

The non-Hermitian skin effect has so far been explored as a means for squeezing bulk states towards the edges of a system. Recently however, Wang et al. reported an opposite process through which the localized topological modes can morph into more general diverse profiles. The resulting extended modes are still robust against local perturbations and can be reshaped by manipulating the non-Hermiticity distribution [117]. In a paradigm shift study, it has been recently shown that the gain/loss induced second order hybrid skin topological modes in two dimensional Chern insulators can be realized without imposing asymmetric hoppings. If the applied gain/loss does not close the line energy gap associated with these open topological systems, all the edge modes tend to localize at one corner, while the bulk states happen to remain extended. By increasing the gain/loss contrast above a certain threshold, a PT phase transition occurs, and an exceptional point (EP) emerges between pairs of skin topological modes. Interestingly, when the gain/loss contrast exceeds a certain threshold, a PT phase transition occurs, and an exceptional point emerges between pairs of skin topological modes [118,119]. Such platforms may open up new venues for exploring the synergistic operation of PT symmetry and topology in higher dimensional settings. Along similar lines, in the context of Floquet systems, it has been shown that periodic quenching can exhibit a non-Hermitian skin effect [120].

4. Non-Hermitian high-order topological insulators

In an effort to explain the observed corner localization, recent studies have indicated a link between higher- order skin effects and higher-order topological insulator phases (HOTI) [121,122]. Supporting N-2 or less protected topological modes, such that the bulk-edge correspondence is extended to a bulk-hinge or bulk-corner correspondence, is a characteristic feature of HOTIs. For instance, 2D (3D) structures manifesting second-order topology exhibit gapped surfaces and edges (gapped bulks and surfaces) and can support topologically protected gapless states at the intersection of two edges (surfaces) – the so called corner (hinge) states. It turns out that the dimensionality of these states can be, indeed, determined by the quantized electric multipole moments of the bulk [123125]. HOTIs in Hermitian systems have been lately a subject of active research [126129] while their non-Hermitian counterparts are only now being investigated. Introducing staggered on-site gain and loss, for example, has been shown to drive a trivial system into the topological domain and promote four degenerate corner states. Non-Hermitian topological invariants, built around nested Wilson loops for this framework, have been found to be protected by reflection or chiral symmetry [130]. Implementing this arrangement in an acoustic crystal suggests that the same behavior can be achieved simply by carefully designing the loss configuration [131]. Along these lines, non-Hermitian second-order topological insulators (SOTIs) are reported in spite of first order topologically trivial bulk bands. Given that the Hamiltonian of a two-dimensional case respects mirror-rotation and sublattice symmetry, the associated midgap robust states are only located at one corner. In three-dimensions, second-order topological modes are anomalously localized at a corner and not along the hinges–signifying the breakdown of the usual bulk-hinge correspondence in the presence of non-Hermiticity. If the underlying twofold mirror-rotation symmetry breaks, then, the modes are localized at more than one corner [132]. A concept unique to the non-Hermitian regime is the clock symmetry that was found to induce topological corner states when imposed on a breathing Kagome lattice [33]. Another example is the hosting of Wannier-type corner states by photonic graphene with mirror-symmetrical gain and loss patterns [133]. These tight binding predictions are still waiting for an experimental verification. On a different front, Yu et al. have explored the zero energy modes in a non-Hermitian quadrupole insulator using a Jordan decomposition. This study reveals the importance of replacing the conventional eigen decomposition with a Jordan decomposition when analyzing the response of non-Hermitian frameworks [134].

5. Topological features of non-Hermitian singularities

Exceptional points (EPs) are known to be among the most prominent and intriguing features of non-Hermitian systems. At an EP not only two or more eigenvalues tend to coalesce but also their corresponding eigenvectors simultaneously collapse on each other–thus leading to an abrupt reduction in dimensionality, an aspect that has no counterpart in a Hermitian environment [135144]. These peculiar band singularities are characterized by a gap closing and as such they play a pivotal role in a rich array of anomalous topological phenomena [97,145,146]. For instance, by exposing a two-dimensional periodic photonic crystal slab to radiative losses, a pair of EPs emerges from a single Dirac point (DP) with a nontrivial Berry phase [147,148]. The associated complex band structure is endowed with a bulk Fermi arc while the polarization of the far field emission signifies the signature of a half integer topological index associated with EPs–both being byproducts of the non-Hermitian topological properties of exceptional points (Fig. 3(a)) [148]. A similar idea has been exploited in Weyl arrangements where a dissipative perturbation transforms isolated point degeneracies (i.e., Weyl points) into Weyl exceptional rings. Contrary to Weyl points and Weyl nodal rings, they possess a nontrivial Chern number and a quantized Berry phase [149151]. Experimental demonstration of this effect provided the first evidence of a distributed Berry flux and demonstrated that the Berry charge and Fermi arc surface states survive in the presence of non-Hermiticity (Fig. 3(b)) [151]. This aspect is further enhanced by the appearance of a Weyl ring in thermal diffusive systems that could be potentially used in topological heat transfer [152].

 figure: Fig. 3.

Fig. 3. Topological features of non-Hermitian degeneracies or exceptional points. (a). A topologically protected bulk Fermi arc that connects two exceptional points emerges from a single Dirac point in a 2D-periodic photonic crystal when subjected to radiative losses [148]. (b). Weyl exceptional ring observed in the band structure of a bipartite helical waveguide array with tunable losses [151]. (c) Topological lasing in a coupled waveguide demonstrating dynamic encirclement of an EP [168]. (d) Monitoring the temporal evolution of light pulse during the parametric steering of a non-Hermitian Hamiltonian [169].

Download Full Size | PDF

On a different front, the distinctive topology of self-intersecting Riemann manifolds around EPs leads to a remarkably different manifestation of the adiabatic theorem in non-Hermitian arrangements. Slow and cyclic transformations around Hermitian degeneracies return the system to itself apart from acquiring a geometric Berry phase. However, when non-Hermitian degeneracies are involved, two possible scenarios are now possible. If a non-Hermitian Hamiltonian steers around an exceptional point in a quasi-static fashion, the eigenvalues of a two-state system tend to swap with each other at the end of the loop [153,154]. In the course of such a process, one of the eigenstates picks up a topological phase shift of $\pi $. Evidently, this requires four encirclement rounds around an EP in order to retrieve the initial state. This peculiar behavior has been observed in two-state exciton-polariton systems [155] and also in a microwave billiard arrangement [156]. However, when an EP is dynamically encircled, non-hermiticity and the ensuing non-adiabatic transitions result in a counterintuitive chiral state transfer. In this regard, for any arbitrary input that happens to be a mixture of the two eigenstates, after winding around an EP in a clockwise (CW) fashion, the system generates in a robust and faithful manner the fundamental mode $|1\rangle$, while in the opposite direction, the output always happens to be the second mode $|2\rangle$, irrespective of the input state.

The dynamics of such systems are governed by the time dependent Schrödinger equation $i{\partial _t}|{\psi (t )\rangle + H(t )} |\psi (t )\rangle = 0$, where their associated Hamiltonian $H(t )$ is given by

$$H(t )= \left( {\begin{array}{cc} { - i\tilde{g}(t )- \tilde{\delta }(t )}&\kappa \\ \kappa &{i\tilde{g}(t )+ \tilde{\delta }(t )} \end{array}} \right)$$
and the state vector is represented by $|\psi (t )\rangle $. The time-varying parameters $\tilde{g}(t )$ and $\tilde{\delta }(t )$ denote the gain/loss and detuning between the two coupled entities respectively and $\kappa $ stands for the coupling strength [157159]. Subsequently, this striking result was analytically explained using the properties of confluent hypergeometric functions [160,161]. The chiral state transfer and Berry phase accumulation in such non-Hermitian frameworks are topologically protected and rely solely on the adiabatic encircling of an EP degeneracy. These fascinating processes were found to be resilient against perturbations that could affect the encirclement trajectory in the parameter space. As such they have sparked an array of experiments in both the optical and the microwave regime [162166]. Observation of topological state transfer along all directions in k space has been used for manifesting rings of exceptional points [167]. As opposed to previous reports in other passive systems, exploring topological mode transfer in active settings allows one to observe behaviors in the entire phase space, without being restricted by severe losses that tend to obscure the output in a significant domain of the operating parameters. Leveraging this effect for lasing action, not only has proved its robustness in the nonlinear domain but also expanded the family of topological lasers (Fig. 3(c)) [168]. The first temporal observation of state evolution during the parametric steering of a non-Hermitian Hamiltonian has been carried in an active fiber-based photonic platform (Fig. 3(d)) [169]. Going beyond the limit of a two-level system, the eigenstate exchange in arrangements supporting multiple EPs has been explored by invoking the topological notion of homotopy [170]. The implications of non-Hermitian topology for practical applications were further explored by observing dynamically tunable topological light steering in non-Hermitian lattices featuring an EP [171].

6. Conclusion

Over the years, research at the interface of topology and non-Hermiticity has been enriched by a series of breakthroughs in these two fields. While topological physics had its birth within the framework of quantum mechanics, it is in optics that it found a reincarnation within the realm of non-Hermitian systems. Non-Hermiticity allows for the emergence of a new class of symmetries, ramifying the existent symmetry classifications substantially – an aspect enriching topological phases beyond their Hermitian counterparts. The interplay of non-Hermiticity with topology and nonlinearity in lasing schemes, has changed the common mindset about fundamental features of topological systems, thus generating a novel and robust class of lasers. The unique topological aspects of non-Hermitian degeneracies have opened another route to design photonic systems that are resilient against operational degradation, fabrication imperfection and environmental perturbations.

Yet, on most occasions, elements in optical systems tend to couple only to nearest neighbors. As such, new developments in realizing long range interactions and in employing synthetic dimensions of higher order are expected to dramatically accelerate the pace of research in this field [56,172175]. Coupled waveguides [173], fiber loops [8,175], and integrated ring resonators [176] are amongst various promising platforms that have been successfully harnessed so far to realize synthetic dimensions. From a scientific standpoint, a universal and self-consistent theory of non-Hermitian bulk-edge-correspondence is still elusive. Another area of active research deals with the ramifications of non-Hermitian topology in open quantum systems. The prospect for topological phases in aperiodic media calls for the development of local and real-space definitions of topology [177,178]. It still remains to be seen whether new non-Hermitian symmetries can be classified as extended members within the broader topological family of insulators. Identifying non-Hermitian topological phases in the presence of many-body interaction is another emerging topic [179,180]. Non-Hermitian topological photonics also holds promise for manipulating the emission pattern of plasmonic nanoantennas, Purcell enhancement enabled by topological edge state, and developing topological beam emitters [181,182]. At a fundamental level, understanding various topological phases that are unique to Floquet and non-Hermitian arrangements is of immediate significance. In light of the recent progress, one can look forward to new research initiatives and paradigms in the field that could find their own niche applications for the next few years to come.

Funding

Air Force Office of Scientific Research (FA9550-20-1-0322, FA9550-21-1-0202); Office of Naval Research (N00014-19-1- 2052, N00014-20-1-2522, N00014-20-1-2789); Defense Advanced Research Projects Agency (D18AP00058); National Science Foundation (CBET 1805200); W. M. Keck Foundation; United States-Israel Binational Science Foundation (BSF: 2016381); Simons Foundation (Simons grant 733682); Air Force Research Laboratory (FA86511820019); Qatar National Research Fund (grant NPRP13S0121-200126).

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References

1. C. M. Bender and S. Boettcher, “Real spectra in non-Hermitian Hamiltonians having PT symmetry,” Phys. Rev. Lett. 80(24), 5243–5246 (1998). [CrossRef]  

2. R. El-Ganainy, K. G. Makris, D. N. Christodoulides, and Z. H. Musslimani, “Theory of coupled optical PT-symmetric structures,” Opt. Lett. 32(17), 2632 (2007). [CrossRef]  

3. K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, “Beam dynamics in PT symmetric optical lattices,” Phys. Rev. Lett. 100(10), 103904 (2008). [CrossRef]  

4. S. Klaiman, U. Günther, and N. Moiseyev, “Visualization of branch points in PT-symmetric waveguides,” Phys. Rev. Lett. 101(8), 080402 (2008). [CrossRef]  

5. H. Ramezani, T. Kottos, R. El-Ganainy, and D. N. Christodoulides, “Unidirectional nonlinear PT-symmetric optical structures,” Phys. Rev. A 82(4), 043803 (2010). [CrossRef]  

6. A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. A. Siviloglou, and D. N. Christodoulides, “Observation of PT-symmetry breaking in complex optical potentials,” Phys. Rev. Lett. 103(9), 093902 (2009). [CrossRef]  

7. C. E. Rüter, K. G. Makris, R. El-Ganainy, D. N. Christodoulides, M. Segev, and D. Kip, “Observation of parity–time symmetry in optics,” Nat. Phys. 6(3), 192–195 (2010). [CrossRef]  

8. A. Regensburger, C. Bersch, M.-A. Miri, G. Onishchukov, D. N. Christodoulides, and U. Peschel, “Parity–time synthetic photonic lattices,” Nature 488(7410), 167–171 (2012). [CrossRef]  

9. R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, “Non-Hermitian physics and PT symmetry,” Nat. Phys. 14(1), 11–19 (2018). [CrossRef]  

10. Z. H. Musslimani, K. G. Makris, R. El-Ganainy, and D. N. Christodoulides, “Optical solitons in PT periodic potentials,” Phys. Rev. Lett. 100(3), 030402 (2008). [CrossRef]  

11. Y. Ashida, Z. Gong, and M. Ueda, “Non-Hermitian physics,” Adv. Phys. 69(3), 249–435 (2020). [CrossRef]  

12. Ş. K. Özdemir, S. Rotter, F. Nori, and L. Yang, “Parity–time symmetry and exceptional points in photonics,” Nat. Mater. 18(8), 783–798 (2019). [CrossRef]  

13. D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, “Quantized hall conductance in a two-dimensional periodic potential,” Phys. Rev. Lett. 49(6), 405–408 (1982). [CrossRef]  

14. D. R. Nelson and J. M. Kosterlitz, “Universal jump in the superfluid density of two-dimensional superfluids,” Phys. Rev. Lett. 39(19), 1201–1205 (1977). [CrossRef]  

15. K. v. Klitzing, G. Dorda, and M. Pepper, “New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance,” Phys. Rev. Lett. 45(6), 494–497 (1980). [CrossRef]  

16. F. D. M. Haldane, “Nonlinear field theory of large-spin heisenberg antiferromagnets: semiclassically quantized solitons of the one-dimensional easy-axis Néel state,” Phys. Rev. Lett. 50(15), 1153–1156 (1983). [CrossRef]  

17. F. D. M. Haldane, ““Model for a quantum hall effect without Landau levels: condensed-matter realization of the “parity anomaly,”,” Phys. Rev. Lett. 61(18), 2015–2018 (1988). [CrossRef]  

18. H. Price, Y. Chong, and A. Khanikaev, “Roadmap on topological photonics,” JPhys Photonics 4(3), 032501 (2022). [CrossRef]  

19. M. Segev and M. A. Bandres, “Topological photonics: Where do we go from here?” Nanophotonics 10(1), 425–434 (2020). [CrossRef]  

20. T. Ozawa, H. M. Price, A. Amo, N. Goldman, M. Hafezi, L. Lu, M. C. Rechtsman, D. Schuster, J. Simon, O. Zilberberg, and I. Carusotto, “Topological photonics,” Rev. Mod. Phys. 91(1), 015006 (2019). [CrossRef]  

21. J. M. Zeuner, M. C. Rechtsman, Y. Plotnik, Y. Lumer, S. Nolte, M. S. Rudner, M. Segev, and A. Szameit, “Observation of a topological transition in the bulk of a non-Hermitian system,” Phys. Rev. Lett. 115(4), 040402 (2015). [CrossRef]  

22. M. Z. Hasan and C. L. Kane, “Colloquium : Topological insulators,” Rev. Mod. Phys. 82(4), 3045–3067 (2010). [CrossRef]  

23. C.-K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, “Classification of topological quantum matter with symmetries,” Rev. Mod. Phys. 88(3), 035005 (2016). [CrossRef]  

24. K. Kawabata, K. Shiozaki, M. Ueda, and M. Sato, “Symmetry and topology in non-Hermitian physics,” Phys. Rev. X 9(4), 041015 (2019). [CrossRef]  

25. Z. Gong, Y. Ashida, K. Kawabata, K. Takasan, S. Higashikawa, and M. Ueda, “Topological phases of non-Hermitian systems,” Phys. Rev. X 8(3), 031079 (2018). [CrossRef]  

26. M. P. Hokmabadi, N. S. Nye, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, “Supersymmetric laser arrays,” Science 363(6427), 623–626 (2019). [CrossRef]  

27. C.-H. Liu, H. Jiang, and S. Chen, “Topological classification of non-Hermitian systems with reflection symmetry,” Phys. Rev. B 99(12), 125103 (2019). [CrossRef]  

28. H. Zhou and J. Y. Lee, “Periodic table for topological bands with non-Hermitian symmetries,” Phys. Rev. B 99(23), 235112 (2019). [CrossRef]  

29. R. El-Ganainy, L. Ge, M. Khajavikhan, and D. N. Christodoulides, “Supersymmetric laser arrays,” Phys. Rev. A 92(3), 033818 (2015). [CrossRef]  

30. K. Kawabata, S. Higashikawa, Z. Gong, Y. Ashida, and M. Ueda, “Topological unification of time-reversal and particle-hole symmetries in non-Hermitian physics,” Nat. Commun. 10(1), 297 (2019). [CrossRef]  

31. K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, “PT -symmetric optical lattices,” Phys. Rev. A 81(6), 063807 (2010). [CrossRef]  

32. R. El-Ganainy, M. Khajavikhan, D. N. Christodoulides, and S. K. Ozdemir, “The dawn of non-Hermitian optics,” Commun. Phys. 2(1), 37 (2019). [CrossRef]  

33. M. Ezawa, “Clock-symmetric non-Hermitian second-order topological insulator,” Eur. Phys. J. B 95(6), 97 (2022). [CrossRef]  

34. H. Hodaei, M.-A. Miri, M. Heinrich, D. N. Christodoulides, and M. Khajavikhan, “Parity-time–symmetric microring lasers,” Science 346(6212), 975–978 (2014). [CrossRef]  

35. H. Hodaei, M. Miri, A. U. Hassan, W. E. Hayenga, M. Heinrich, D. N. Christodoulides, and M. Khajavikhan, “Single mode lasing in transversely multi-moded PT-symmetric microring resonators,” Laser Photonics Rev. 10(3), 494–499 (2016). [CrossRef]  

36. C. Xu, W. E. Hayenga, H. Hodaei, D. N. Christodoulides, M. Khajavikhan, and P. LiKamWa, “Enhanced modulation characteristics in broken symmetric coupled microring lasers,” Opt. Express 28(13), 19608 (2020). [CrossRef]  

37. W. B. Rui, Z. Zheng, C. Wang, and Z. D. Wang, “Non-Hermitian spatial symmetries and their stabilized normal and exceptional topological semimetals,” Phys. Rev. Lett. 128(22), 226401 (2022). [CrossRef]  

38. C.-H. Liu, H. Hu, and S. Chen, “Symmetry and topological classification of Floquet non-Hermitian systems,” Phys. Rev. B 105(21), 214305 (2022). [CrossRef]  

39. C.-H. Liu and S. Chen, “Topological classification of defects in non-Hermitian systems,” Phys. Rev. B 100(14), 144106 (2019). [CrossRef]  

40. Y. C. Hu and T. L. Hughes, “Absence of topological insulator phases in non-Hermitian PT-symmetric Hamiltonians,” Phys. Rev. B 84(15), 153101 (2011). [CrossRef]  

41. C. Yuce, “PT symmetric Floquet topological phase,” Eur. Phys. J. D 69(7), 184 (2015). [CrossRef]  

42. C. Yuce, “Topological phase in a non-Hermitian PT symmetric system,” Phys. Lett. A 379(18-19), 1213–1218 (2015). [CrossRef]  

43. A. K. Harter, T. E. Lee, and Y. N. Joglekar, “PT -breaking threshold in spatially asymmetric Aubry-André and Harper models: hidden symmetry and topological states,” Phys. Rev. A 93(6), 062101 (2016). [CrossRef]  

44. S. Weimann, M. Kremer, Y. Plotnik, Y. Lumer, S. Nolte, K. G. Makris, M. Segev, M. C. Rechtsman, and A. Szameit, “Topologically protected bound states in photonic parity–time-symmetric crystals,” Nat. Mater. 16(4), 433–438 (2017). [CrossRef]  

45. C. Poli, M. Bellec, U. Kuhl, F. Mortessagne, and H. Schomerus, “Selective enhancement of topologically induced interface states in a dielectric resonator chain,” Nat. Commun. 6(1), 6710 (2015). [CrossRef]  

46. S. Malzard, C. Poli, and H. Schomerus, “Topologically protected defect states in open photonic systems with non-Hermitian charge-conjugation and parity-time symmetry,” Phys. Rev. Lett. 115(20), 200402 (2015). [CrossRef]  

47. K. Takata and M. Notomi, “Photonic topological insulating phase induced solely by gain and loss,” Phys. Rev. Lett. 121(21), 213902 (2018). [CrossRef]  

48. B. Zhu, L.-J. Lang, Q. Wang, Q. J. Wang, and Y. D. Chong, “Topological transitions with an imaginary Aubry-Andre-Harper potential,” arXiv, arXiv:2211.07853 (2022). [CrossRef]  

49. S. Weidemann, M. Kremer, S. Longhi, and A. Szameit, “Topological triple phase transition in non-Hermitian Floquet quasicrystals,” Nature 601(7893), 354–359 (2022). [CrossRef]  

50. Z. Wang, Y. Chong, J. D. Joannopoulos, and M. Soljačić, “Observation of unidirectional backscattering-immune topological electromagnetic states,” Nature 461(7265), 772–775 (2009). [CrossRef]  

51. Z. Wang, Y. D. Chong, J. D. Joannopoulos, and M. Soljačić, “Reflection-free one-way edge modes in a gyromagnetic photonic crystal,” Phys. Rev. Lett. 100(1), 013905 (2008). [CrossRef]  

52. L. He, Z. Addison, J. Jin, E. J. Mele, S. G. Johnson, and B. Zhen, “Floquet Chern insulators of light,” Nat. Commun. 10(1), 4194 (2019). [CrossRef]  

53. L. J. Maczewsky, J. M. Zeuner, S. Nolte, and A. Szameit, “Observation of photonic anomalous Floquet topological insulators,” Nat. Commun. 8(1), 13756 (2017). [CrossRef]  

54. M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit, “Photonic Floquet topological insulators,” Nature 496(7444), 196–200 (2013). [CrossRef]  

55. G. G. Pyrialakos, J. Beck, M. Heinrich, L. J. Maczewsky, N. V. Kantartzis, M. Khajavikhan, A. Szameit, and D. N. Christodoulides, “Bimorphic Floquet topological insulators,” Nat. Mater. 21(6), 634–639 (2022). [CrossRef]  

56. Y. G. N. Liu, P. S. Jung, M. Parto, D. N. Christodoulides, and M. Khajavikhan, “Gain-induced topological response via tailored long-range interactions,” Nat. Phys. 17(6), 704–709 (2021). [CrossRef]  

57. Y. G. N. Liu, O. Hemmatyar, A. U. Hassan, P. S. Jung, J.-H. Choi, D. N. Christodoulides, and M. Khajavikhan, “Engineering interaction dynamics in active resonant photonic structures,” APL Photonics 6(5), 050804 (2021). [CrossRef]  

58. H. Schomerus, “Topologically protected midgap states in complex photonic lattices,” Opt. Lett. 38(11), 1912 (2013). [CrossRef]  

59. P. St-Jean, V. Goblot, E. Galopin, A. Lemaître, T. Ozawa, L. Le Gratiet, I. Sagnes, J. Bloch, and A. Amo, “Lasing in topological edge states of a one-dimensional lattice,” Nat. Photonics 11(10), 651–656 (2017). [CrossRef]  

60. M. Parto, S. Wittek, H. Hodaei, G. Harari, M. A. Bandres, J. Ren, M. C. Rechtsman, M. Segev, D. N. Christodoulides, and M. Khajavikhan, “Edge-Mode Lasing in 1D Topological Active Arrays,” Phys. Rev. Lett. 120(11), 113901 (2018). [CrossRef]  

61. H. Zhao, P. Miao, M. H. Teimourpour, S. Malzard, R. El-Ganainy, H. Schomerus, and L. Feng, “Topological hybrid silicon microlasers,” Nat. Commun. 9(1), 981 (2018). [CrossRef]  

62. Y. Ota, R. Katsumi, K. Watanabe, S. Iwamoto, and Y. Arakawa, “Topological photonic crystal nanocavity laser,” Commun. Phys. 1(1), 86 (2018). [CrossRef]  

63. C. Han, M. Lee, S. Callard, C. Seassal, and H. Jeon, “Lasing at topological edge states in a photonic crystal L3 nanocavity dimer array,” Light: Sci. Appl. 8(1), 40 (2019). [CrossRef]  

64. M. A. Bandres, S. Wittek, G. Harari, M. Parto, J. Ren, M. Segev, D. N. Christodoulides, and M. Khajavikhan, “Topological insulator laser: Experiments,” Science 359(6381), eaar4005 (2018). [CrossRef]  

65. G. Harari, M. A. Bandres, Y. Lumer, M. C. Rechtsman, Y. D. Chong, M. Khajavikhan, D. N. Christodoulides, and M. Segev, “Topological insulator laser: Theory,” Science 359(6381), eaar4003 (2018). [CrossRef]  

66. M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. M. Taylor, “Imaging topological edge states in silicon photonics,” Nat. Photonics 7(12), 1001–1005 (2013). [CrossRef]  

67. M. Hafezi, E. A. Demler, M. D. Lukin, and J. M. Taylor, “Robust optical delay lines with topological protection,” Nat. Phys. 7(11), 907–912 (2011). [CrossRef]  

68. B. Bahari, A. Ndao, F. Vallini, A. El Amili, Y. Fainman, and B. Kanté, “Nonreciprocal lasing in topological cavities of arbitrary geometries,” Science 358(6363), 636–640 (2017). [CrossRef]  

69. B. Bahari, L. Hsu, S. H. Pan, D. Preece, A. Ndao, A. El Amili, Y. Fainman, and B. Kanté, “Photonic quantum Hall effect and multiplexed light sources of large orbital angular momenta,” Nat. Phys. 17(6), 700–703 (2021). [CrossRef]  

70. S. Klembt, T. H. Harder, O. A. Egorov, K. Winkler, R. Ge, M. A. Bandres, M. Emmerling, L. Worschech, T. C. H. Liew, M. Segev, C. Schneider, and S. Höfling, “Exciton-polariton topological insulator,” Nature 562(7728), 552–556 (2018). [CrossRef]  

71. I. Amelio and I. Carusotto, “Theory of the coherence of topological lasers,” Phys. Rev. X 10(4), 041060 (2020). [CrossRef]  

72. A. Dikopoltsev, T. H. Harder, E. Lustig, O. A. Egorov, J. Beierlein, A. Wolf, Y. Lumer, M. Emmerling, C. Schneider, S. Höfling, M. Segev, and S. Klembt, “Topological insulator vertical-cavity laser array,” Science 373(6562), 1514–1517 (2021). [CrossRef]  

73. Z. Yang, E. Lustig, G. Harari, Y. Plotnik, Y. Lumer, M. A. Bandres, and M. Segev, “Mode-Locked topological insulator laser utilizing synthetic dimensions,” Phys. Rev. X 10(1), 011059 (2020). [CrossRef]  

74. W. Noh, H. Nasari, H.-M. Kim, Q. Le-Van, Z. Jia, C.-H. Huang, and B. Kanté, “Experimental demonstration of single-mode topological valley-Hall lasing at telecommunication wavelength controlled by the degree of asymmetry,” Opt. Lett. 45(15), 4108 (2020). [CrossRef]  

75. Z.-K. Shao, H.-Z. Chen, S. Wang, X.-R. Mao, Z.-Q. Yang, S.-L. Wang, X.-X. Wang, X. Hu, and R.-M. Ma, “A high-performance topological bulk laser based on band-inversion-induced reflection,” Nat. Nanotechnol. 15(1), 67–72 (2020). [CrossRef]  

76. Z.-Q. Yang, Z.-K. Shao, H.-Z. Chen, X.-R. Mao, and R.-M. Ma, “Spin-momentum-locked edge mode for topological vortex lasing,” Phys. Rev. Lett. 125(1), 013903 (2020). [CrossRef]  

77. M. Ezawa, “Nonlinear non-Hermitian higher-order topological laser,” Phys. Rev. Res. 4(1), 013195 (2022). [CrossRef]  

78. H. Zhong, Y. V. Kartashov, A. Szameit, Y. Li, C. Liu, and Y. Zhang, “Theory of topological corner state laser in Kagome waveguide arrays,” APL Photonics 6(4), 040802 (2021). [CrossRef]  

79. W. Zhang, X. Xie, H. Hao, J. Dang, S. Xiao, S. Shi, H. Ni, Z. Niu, C. Wang, K. Jin, X. Zhang, and X. Xu, “Low-threshold topological nanolasers based on the second-order corner state,” Light: Sci. Appl. 9(1), 109 (2020). [CrossRef]  

80. C. Han, M. Kang, and H. Jeon, “Lasing at multidimensional topological states in a two-dimensional photonic crystal structure,” ACS Photonics 7(8), 2027–2036 (2020). [CrossRef]  

81. H.-R. Kim, M.-S. Hwang, D. Smirnova, K.-Y. Jeong, Y. Kivshar, and H.-G. Park, “Multipolar lasing modes from topological corner states,” Nat. Commun. 11(1), 5758 (2020). [CrossRef]  

82. X. Gao, L. Yang, H. Lin, L. Zhang, J. Li, F. Bo, Z. Wang, and L. Lu, “Dirac-vortex topological cavities,” Nat. Nanotechnol. 15(12), 1012–1018 (2020). [CrossRef]  

83. J. Ma, T. Zhou, M. Tang, H. Li, Z. Zhang, X. Xi, M. Martin, T. Baron, H. Liu, Z. Zhang, S. Chen, and X. Sun, “Room-temperature continuous-wave Dirac-vortex topological lasers on silicon,” arXiv, arXiv:2106.13838 (2021).

84. L. Yang, G. Li, X. Gao, and L. Lu, “Topological-cavity surface-emitting laser,” Nat. Photonics 16(4), 279–283 (2022). [CrossRef]  

85. B. Bahari, J.-H. Choi, Y. G. N. Liu, and M. Khajavikhan, “Majorana bound state cavities,” Phys. Rev. B 104(23), 235423 (2021). [CrossRef]  

86. Y. Zeng, U. Chattopadhyay, B. Zhu, B. Qiang, J. Li, Y. Jin, L. Li, A. G. Davies, E. H. Linfield, B. Zhang, Y. Chong, and Q. J. Wang, “Electrically pumped topological laser with valley edge modes,” Nature 578(7794), 246–250 (2020). [CrossRef]  

87. J.-H. Choi, W. E. Hayenga, Y. G. N. Liu, M. Parto, B. Bahari, D. N. Christodoulides, and M. Khajavikhan, “Room temperature electrically pumped topological insulator lasers,” Nat. Commun. 12(1), 3434 (2021). [CrossRef]  

88. C. Leefmans, M. Parto, J. Williams, G. H. Y. Li, A. Dutt, F. Nori, and A. Marandi, “Topological temporally mode-locked laser,” arXiv, arXiv:2209.00762 (2022). [CrossRef]  

89. S. Yao and Z. Wang, “Edge states and topological invariants of non-Hermitian systems,” Phys. Rev. Lett. 121(8), 086803 (2018). [CrossRef]  

90. N. Okuma, K. Kawabata, K. Shiozaki, and M. Sato, “Topological 0rigin of Nnn-Hermitian skin effects,” Phys. Rev. Lett. 124(8), 086801 (2020). [CrossRef]  

91. T. E. Lee, “Anomalous edge state in a non-Hermitian lattice,” Phys. Rev. Lett. 116(13), 133903 (2016). [CrossRef]  

92. K. Kawabata, K. Shiozaki, and M. Ueda, “Anomalous helical edge states in a non-Hermitian Chern insulator,” Phys. Rev. B 98(16), 165148 (2018). [CrossRef]  

93. S. Yao, F. Song, and Z. Wang, “Non-Hermitian Chern bands,” Phys. Rev. Lett. 121(13), 136802 (2018). [CrossRef]  

94. K. Yokomizo and S. Murakami, “Non-Bloch band theory of non-Hermitian systems,” Phys. Rev. Lett. 123(6), 066404 (2019). [CrossRef]  

95. F. K. Kunst, E. Edvardsson, J. C. Budich, and E. J. Bergholtz, “Biorthogonal Bulk-Boundary Correspondence in Non-Hermitian Systems,” Phys. Rev. Lett. 121(2), 026808 (2018). [CrossRef]  

96. D. Nakamura, T. Bessho, and M. Sato, “Bulk-boundary correspondence in point-gap topological phases,” arXiv, arXiv:2205.15635 (2022). [CrossRef]  

97. T. Bessho, K. Kawabata, and M. Sato, “Topological classificaton of non-hermitian gapless phases: exceptional points and bulk Fermi arcs,” in Proceedings of the International Conference on Strongly Correlated Electron Systems (SCES2019) (2020). [CrossRef]  

98. Z. Yang, K. Zhang, C. Fang, and J. Hu, “Non-Hermitian bulk-boundary correspondence and auxiliary generalized Brillouin zone theory,” Phys. Rev. Lett. 125(22), 226402 (2020). [CrossRef]  

99. A. Ghatak and T. Das, “New topological invariants in non-Hermitian systems,” J. Phys.: Condens. Matter 31(26), 263001 (2019). [CrossRef]  

100. P. Bouteyre, D. X. Nguyen, G. Gachon, T. Benyattou, X. Letartre, P. Viktorovitch, S. Callard, L. Ferrier, and H. S. Nguyen, “Non-Hermitian topological invariant of photonic band structures undergoing inversion,” arXiv, arXiv:2211.09884 (2022).

101. N. Hatano and D. R. Nelson, “Localization transitions in non-Hermitian quantum mechanics,” Phys. Rev. Lett. 77(3), 570–573 (1996). [CrossRef]  

102. N. Hatano and D. R. Nelson, “Non-Hermitian delocalization and eigenfunctions,” Phys. Rev. B 58(13), 8384–8390 (1998). [CrossRef]  

103. S. Weidemann, M. Kremer, T. Helbig, T. Hofmann, A. Stegmaier, M. Greiter, R. Thomale, and A. Szameit, “Topological funneling of light,” Science 368(6488), 311–314 (2020). [CrossRef]  

104. L. Xiao, T. Deng, K. Wang, G. Zhu, Z. Wang, W. Yi, and P. Xue, “Observation of non-Hermitian bulk-boundary correspondence in quantum dynamics,” Nat. Phys. 16(7), 761–766 (2020). [CrossRef]  

105. K. Wang, A. Dutt, K. Y. Yang, C. C. Wojcik, J. Vučković, and S. Fan, “Generating arbitrary topological windings of a non-Hermitian band,” Science 371(6535), 1240–1245 (2021). [CrossRef]  

106. Y. G. N. Liu, Y. Wei, O. Hemmatyar, G. G. Pyrialakos, P. S. Jung, D. N. Christodoulides, and M. Khajavikhan, “Complex skin modes in non-Hermitian coupled laser arrays,” Light: Sci. Appl. 11(1), 336 (2022). [CrossRef]  

107. A. McDonald and A. A. Clerk, “Exponentially-enhanced quantum sensing with non-Hermitian lattice dynamics,” Nat. Commun. 11(1), 5382 (2020). [CrossRef]  

108. C. H. Lee and S. Longhi, “Ultrafast and anharmonic Rabi oscillations between non-Bloch bands,” Commun. Phys. 3(1), 147 (2020). [CrossRef]  

109. J. C. Budich and E. J. Bergholtz, “Non-Hermitian Topological Sensors,” Phys. Rev. Lett. 125(18), 180403 (2020). [CrossRef]  

110. D. Zou, T. Chen, W. He, J. Bao, C. H. Lee, H. Sun, and X. Zhang, “Observation of hybrid higher-order skin-topological effect in non-Hermitian topolectrical circuits,” Nat. Commun. 12(1), 7201 (2021). [CrossRef]  

111. L. Zhang, Y. Yang, Y. Ge, Y.-J. Guan, Q. Chen, Q. Yan, F. Chen, R. Xi, Y. Li, D. Jia, S.-Q. Yuan, H.-X. Sun, H. Chen, and B. Zhang, “Acoustic non-Hermitian skin effect from twisted winding topology,” Nat. Commun. 12(1), 6297 (2021). [CrossRef]  

112. X. Zhang, Y. Tian, J.-H. Jiang, M.-H. Lu, and Y.-F. Chen, “Observation of higher-order non-Hermitian skin effect,” Nat. Commun. 12(1), 5377 (2021). [CrossRef]  

113. A. Ghatak, M. Brandenbourger, J. van Wezel, and C. Coulais, “Observation of non-Hermitian topology and its bulk–edge correspondence in an active mechanical metamaterial,” Proc. Natl. Acad. Sci. 117(47), 29561–29568 (2020). [CrossRef]  

114. Q. Liang, D. Xie, Z. Dong, H. Li, H. Li, B. Gadway, W. Yi, and B. Yan, “Observation of non-Hermitian skin effect and topology in ultracold atoms,” Phys. Rev. Lett. 129(7), 070401 (2022). [CrossRef]  

115. S. M. Rafi-Ul-Islam, Z. B. Siu, H. Sahin, C. H. Lee, and M. B. A. Jalil, “Unconventional skin modes in generalized topolectrical circuits with multiple asymmetric couplings,” Phys. Rev. Res. 4(4), 043108 (2022). [CrossRef]  

116. W.-T. Xue, Y.-M. Hu, F. Song, and Z. Wang, “Non-Hermitian edge burst,” Phys. Rev. Lett. 128(12), 120401 (2022). [CrossRef]  

117. W. Wang, X. Wang, and G. Ma, “Non-Hermitian morphing of topological modes,” Nature 608(7921), 50–55 (2022). [CrossRef]  

118. W. Zhu and J. Gong, “Hybrid skin-topological modes without asymmetric couplings,” Phys. Rev. B 106(3), 035425 (2022). [CrossRef]  

119. Y. Li, C. Liang, C. Wang, C. Lu, and Y.-C. Liu, “Gain-loss-induced hybrid skin-topological effect,” Phys. Rev. Lett. 128(22), 223903 (2022). [CrossRef]  

120. X. Zhang and J. Gong, “Non-Hermitian Floquet topological phases: Exceptional points, coalescent edge modes, and the skin effect,” Phys. Rev. B 101(4), 045415 (2020). [CrossRef]  

121. K. Kawabata, M. Sato, and K. Shiozaki, “Higher-order non-Hermitian skin effect,” Phys. Rev. B 102(20), 205118 (2020). [CrossRef]  

122. R. Okugawa, R. Takahashi, and K. Yokomizo, “Second-order topological non-Hermitian skin effects,” Phys. Rev. B 102(24), 241202 (2020). [CrossRef]  

123. W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, “Electric multipole moments, topological multipole moment pumping, and chiral hinge states in crystalline insulators,” Phys. Rev. B 96(24), 245115 (2017). [CrossRef]  

124. W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, “Quantized electric multipole insulators,” Science 357(6346), 61–66 (2017). [CrossRef]  

125. F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. P. Parkin, B. A. Bernevig, and T. Neupert, “Higher-order topological insulators,” Sci. Adv. 4(6), 1 (2018). [CrossRef]  

126. S. A. Parameswaran and Y. Wan, “Topological insulators turn a corner,” Physics 10, 132 (2017). [CrossRef]  

127. C. W. Peterson, W. A. Benalcazar, T. L. Hughes, and G. Bahl, “A quantized microwave quadrupole insulator with topologically protected corner states,” Nature 555(7696), 346–350 (2018). [CrossRef]  

128. J. Noh, W. A. Benalcazar, S. Huang, M. J. Collins, K. P. Chen, T. L. Hughes, and M. C. Rechtsman, “Topological protection of photonic mid-gap defect modes,” Nat. Photonics 12(7), 408–415 (2018). [CrossRef]  

129. M. Kim, Z. Jacob, and J. Rho, “Recent advances in 2D, 3D and higher-order topological photonics,” Light: Sci. Appl. 9(1), 130 (2020). [CrossRef]  

130. X.-W. Luo and C. Zhang, “Higher-order topological corner states induced by gain and loss,” Phys. Rev. Lett. 123(7), 073601 (2019). [CrossRef]  

131. H. Gao, H. Xue, Z. Gu, T. Liu, J. Zhu, and B. Zhang, “Non-Hermitian route to higher-order topology in an acoustic crystal,” Nat. Commun. 12(1), 1888 (2021). [CrossRef]  

132. T. Liu, Y.-R. Zhang, Q. Ai, Z. Gong, K. Kawabata, M. Ueda, and F. Nori, “Second-order topological phases in non-Hermitian systems,” Phys. Rev. Lett. 122(7), 076801 (2019). [CrossRef]  

133. Y.-J. Wu, C.-C. Liu, and J. Hou, “Wannier-type photonic higher-order topological corner states induced solely by gain and loss,” Phys. Rev. A 101(4), 043833 (2020). [CrossRef]  

134. Y. Yu, M. Jung, and G. Shvets, “Zero-energy corner states in a non-Hermitian quadrupole insulator,” Phys. Rev. B 103(4), L041102 (2021). [CrossRef]  

135. Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, “Unidirectional Invisibility Induced by P T -Symmetric Periodic Structures,” Phys. Rev. Lett. 106(21), 213901 (2011). [CrossRef]  

136. J. Ren, H. Hodaei, G. Harari, A. U. Hassan, W. Chow, M. Soltani, D. Christodoulides, and M. Khajavikhan, “Ultrasensitive micro-scale parity-time-symmetric ring laser gyroscope,” Opt. Lett. 42(8), 1556 (2017). [CrossRef]  

137. H. Hodaei, A. U. Hassan, S. Wittek, H. Garcia-Gracia, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, “Enhanced sensitivity at higher-order exceptional points,” Nature 548(7666), 187–191 (2017). [CrossRef]  

138. W. E. Hayenga, M. Parto, J. Ren, F. O. Wu, M. P. Hokmabadi, C. Wolff, R. El-Ganainy, N. A. Mortensen, D. N. Christodoulides, and M. Khajavikhan, “Direct generation of tunable orbital angular momentum beams in microring lasers with broadband exceptional points,” ACS Photonics 6(8), 1895–1901 (2019). [CrossRef]  

139. W. Chen, Ş. Kaya Özdemir, G. Zhao, J. Wiersig, and L. Yang, “Exceptional points enhance sensing in an optical microcavity,” Nature 548(7666), 192–196 (2017). [CrossRef]  

140. N. A. Mortensen, P. A. D. Gonçalves, M. Khajavikhan, D. N. Christodoulides, C. Tserkezis, and C. Wolff, “Fluctuations and noise-limited sensing near the exceptional point of parity-time-symmetric resonator systems,” Optica 5(10), 1342 (2018). [CrossRef]  

141. M. P. Hokmabadi, A. Schumer, D. N. Christodoulides, and M. Khajavikhan, “Non-Hermitian ring laser gyroscopes with enhanced Sagnac sensitivity,” Nature 576(7785), 70–74 (2019). [CrossRef]  

142. Q. Zhong, J. Ren, M. Khajavikhan, D. N. Christodoulides, Ş. K. Özdemir, and R. El-Ganainy, “Sensing with exceptional surfaces in order to combine sensitivity with robustness,” Phys. Rev. Lett. 122(15), 153902 (2019). [CrossRef]  

143. H. Hodaei, M. A. Miri, A. U. Hassan, W. E. Hayenga, M. Heinrich, D. N. Christodoulides, and M. Khajavikhan, “Parity-time-symmetric coupled microring lasers operating around an exceptional point,” Opt. Lett. 40(21), 4955 (2015). [CrossRef]  

144. F. Yang, A. Hwang, C. Doiron, and G. V. Naik, “Non-Hermitian metasurfaces for the best of plasmonics and dielectrics,” Opt. Mater. Express 11(7), 2326 (2021). [CrossRef]  

145. M.-A. Miri and A. Alù, “Exceptional points in optics and photonics,” Science 363(6422), eaar7709 (2019). [CrossRef]  

146. M. Parto, Y. G. N. Liu, B. Bahari, M. Khajavikhan, and D. N. Christodoulides, “Non-Hermitian and topological photonics: optics at an exceptional point,” Nanophotonics 10(1), 403–423 (2020). [CrossRef]  

147. Ş. K. Özdemir, “Fermi arcs connect topological degeneracies,” Science 359(6379), 995–996 (2018). [CrossRef]  

148. H. Zhou, C. Peng, Y. Yoon, C. W. Hsu, K. A. Nelson, L. Fu, J. D. Joannopoulos, M. Soljačić, and B. Zhen, “Observation of bulk Fermi arc and polarization half charge from paired exceptional points,” Science 359(6379), 1009–1012 (2018). [CrossRef]  

149. Y. Xu, S.-T. Wang, and L.-M. Duan, “Weyl exceptional rings in a three-dimensional dissipative cold atomic gas,” Phys. Rev. Lett. 118(4), 045701 (2017). [CrossRef]  

150. A. Cerjan, M. Xiao, L. Yuan, and S. Fan, “Effects of non-Hermitian perturbations on Weyl Hamiltonians with arbitrary topological charges,” Phys. Rev. B 97(7), 075128 (2018). [CrossRef]  

151. A. Cerjan, S. Huang, M. Wang, K. P. Chen, Y. Chong, and M. C. Rechtsman, “Experimental realization of a Weyl exceptional ring,” Nat. Photonics 13(9), 623–628 (2019). [CrossRef]  

152. G. Xu, W. Li, X. Zhou, H. Li, Y. Li, S. Fan, S. Zhang, D. N. Christodoulides, and C.-W. Qiu, “Observation of Weyl exceptional rings in thermal diffusion,” Proc. Natl. Acad. Sci. 119(15), e2110018119 (2022). [CrossRef]  

153. W. D. Heiss, “The physics of exceptional points,” J. Phys. Math. Theor. 45(44), 444016 (2012). [CrossRef]  

154. A. A. Mailybaev, O. N. Kirillov, and A. P. Seyranian, “Geometric phase around exceptional points,” Phys. Rev. A 72(1), 014104 (2005). [CrossRef]  

155. T. Gao, E. Estrecho, K. Y. Bliokh, T. C. H. Liew, M. D. Fraser, S. Brodbeck, M. Kamp, C. Schneider, S. Höfling, Y. Yamamoto, F. Nori, Y. S. Kivshar, A. G. Truscott, R. G. Dall, and E. A. Ostrovskaya, “Observation of non-Hermitian degeneracies in a chaotic exciton-polariton billiard,” Nature 526(7574), 554–558 (2015). [CrossRef]  

156. C. Dembowski, H.-D. Gräf, H. L. Harney, A. Heine, W. D. Heiss, H. Rehfeld, and A. Richter, “Experimental observation of the topological structure of exceptional points,” Phys. Rev. Lett. 86(5), 787–790 (2001). [CrossRef]  

157. R. Uzdin, A. Mailybaev, and N. Moiseyev, “On the observability and asymmetry of adiabatic state flips generated by exceptional points,” J. Phys. Math. Theor. 44(43), 435302 (2011). [CrossRef]  

158. E.-M. Graefe, A. A. Mailybaev, and N. Moiseyev, “Breakdown of adiabatic transfer of light in waveguides in the presence of absorption,” Phys. Rev. A 88(3), 033842 (2013). [CrossRef]  

159. I. Gilary, A. A. Mailybaev, and N. Moiseyev, “Time-asymmetric quantum-state-exchange mechanism,” Phys. Rev. A 88(1), 010102 (2013). [CrossRef]  

160. A. U. Hassan, B. Zhen, M. Soljačić, M. Khajavikhan, and D. N. Christodoulides, “Dynamically Encircling Exceptional Points: Exact Evolution and Polarization State Conversion,” Phys. Rev. Lett. 118(9), 093002 (2017). [CrossRef]  

161. A. U. Hassan, G. L. Galmiche, G. Harari, P. LiKamWa, M. Khajavikhan, M. Segev, and D. N. Christodoulides, “Chiral state conversion without encircling an exceptional point,” Phys. Rev. A 96(5), 052129 (2017). [CrossRef]  

162. J. Doppler, A. A. Mailybaev, J. Böhm, U. Kuhl, A. Girschik, F. Libisch, T. J. Milburn, P. Rabl, N. Moiseyev, and S. Rotter, “Dynamically encircling an exceptional point for asymmetric mode switching,” Nature 537(7618), 76–79 (2016). [CrossRef]  

163. H. Xu, D. Mason, L. Jiang, and J. G. E. Harris, “Topological energy transfer in an optomechanical system with exceptional points,” Nature 537(7618), 80–83 (2016). [CrossRef]  

164. J. W. Yoon, Y. Choi, C. Hahn, G. Kim, S. H. Song, K.-Y. Yang, J. Y. Lee, Y. Kim, C. S. Lee, J. K. Shin, H.-S. Lee, and P. Berini, “Time-asymmetric loop around an exceptional point over the full optical communications band,” Nature 562(7725), 86–90 (2018). [CrossRef]  

165. X.-L. Zhang and C. T. Chan, “Dynamically encircling exceptional points in a three-mode waveguide system,” Commun. Phys. 2(1), 63 (2019). [CrossRef]  

166. Y. Choi, J. W. Yoon, J. K. Hong, Y. Ryu, and S. H. Song, “Direct observation of time-asymmetric breakdown of the standard adiabaticity around an exceptional point,” Commun. Phys. 3(1), 140 (2020). [CrossRef]  

167. B. Zhen, C. W. Hsu, Y. Igarashi, L. Lu, I. Kaminer, A. Pick, S.-L. Chua, J. D. Joannopoulos, and M. Soljačić, “Spawning rings of exceptional points out of Dirac cones,” Nature 525(7569), 354–358 (2015). [CrossRef]  

168. A. Schumer, Y. G. N. Liu, J. Leshin, L. Ding, Y. Alahmadi, A. U. Hassan, H. Nasari, S. Rotter, D. N. Christodoulides, P. LiKamWa, and M. Khajavikhan, “Topological modes in a laser cavity through exceptional state transfer,” Science 375(6583), 884–888 (2022). [CrossRef]  

169. H. Nasari, G. Lopez-Galmiche, H. E. Lopez-Aviles, A. Schumer, A. U. Hassan, Q. Zhong, S. Rotter, P. LiKamWa, D. N. Christodoulides, and M. Khajavikhan, “Observation of chiral state transfer without encircling an exceptional point,” Nature 605(7909), 256–261 (2022). [CrossRef]  

170. Q. Zhong, M. Khajavikhan, D. N. Christodoulides, and R. El-Ganainy, “Winding around non-Hermitian singularities,” Nat. Commun. 9(1), 4808 (2018). [CrossRef]  

171. H. Zhao, X. Qiao, T. Wu, B. Midya, S. Longhi, and L. Feng, “Non-Hermitian topological light steering,” Science 365(6458), 1163–1166 (2019). [CrossRef]  

172. E. Lustig and M. Segev, “Topological photonics in synthetic dimensions,” Adv. Opt. Photonics 13(2), 426 (2021). [CrossRef]  

173. E. Lustig, S. Weimann, Y. Plotnik, Y. Lumer, M. A. Bandres, A. Szameit, and M. Segev, “Photonic topological insulator in synthetic dimensions,” Nature 567(7748), 356–360 (2019). [CrossRef]  

174. L. Yuan, Q. Lin, M. Xiao, and S. Fan, “Synthetic dimension in photonics,” Optica 5(11), 1396 (2018). [CrossRef]  

175. A. Dutt, Q. Lin, L. Yuan, M. Minkov, M. Xiao, and S. Fan, “A single photonic cavity with two independent physical synthetic dimensions,” Science 367(6473), 59–64 (2020). [CrossRef]  

176. A. Balčytis, T. Ozawa, Y. Ota, S. Iwamoto, J. Maeda, and T. Baba, “Synthetic dimension band structures on a Si CMOS photonic platform,” Sci. Adv. 8(4), eabk0468 (2022). [CrossRef]  

177. C. Bourne and E. Prodan, “Non-commutative Chern numbers for generic aperiodic discrete systems,” J. Phys. Math. Theor. 51(23), 235202 (2018). [CrossRef]  

178. A. Cerjan and T. A. Loring, “An operator-based approach to topological photonics,” Nanophotonics 11(21), 4765–4780 (2022). [CrossRef]  

179. K. Kawabata, K. Shiozaki, and S. Ryu, “Many-body topology of non-Hermitian systems,” Phys. Rev. B 105(16), 165137 (2022). [CrossRef]  

180. S.-B. Zhang, M. M. Denner, T. Bzdušek, M. A. Sentef, and T. Neupert, “Symmetry breaking and spectral structure of the interacting Hatano-Nelson model,” Phys. Rev. B 106(12), L121102 (2022). [CrossRef]  

181. Z. Qian, Z. Li, H. Hao, L. Shan, Q. Zhang, J. Dong, Q. Gong, and Y. Gu, “Absorption reduction of large Purcell enhancement enabled by topological state-led mode coupling,” Phys. Rev. Lett. 126(2), 023901 (2021). [CrossRef]  

182. K. Y. Lee, S. Yoon, S. H. Song, and J. W. Yoon, “Topological beaming of light,” Sci. Adv. 8(49), eadd8349 (2022). [CrossRef]  

Data availability

No data were generated or analyzed in the presented research.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Topological lasers. (a) Robust lasing in an edge mode in a photonic implementation of the Haldane lattice [56]. Lasing in the protected midgap state of a Su–Schrieffer–Heeger (SSH) lattice implemented using (b) a zigzag chain of coupled polariton micropillars [59] and (c) a microring resonators InGaAsP/InP structure [60]. (d) A topological insulator laser using a two-dimensional array of site and intermediary link resonators that emulates the quantum spin Hall effect [64]. (e) Topological lasing in a quantum Hall platform realized by integrating a photonic crystal with a gyromagnetic substrate [68] (f) Topological insulator vertical-cavity surface-emitting laser (VCSEL) array [72]. (g) Multipolar lasing from corner topological modes in a generalized 2D SSH lattice [81]. (h) An electrically pumped topological insulator laser using a quantum spin Hall platform [87]. (i) topological temporal mode-locking between the temporal modes of a fiber lase cavity [88].
Fig. 2.
Fig. 2. Non-Hermitian skin effect. (a). Efficient funnel for light based on non-Hermitian skin effect in time synthetic photonic mesh lattices [103]. (b). Unidirectional microring laser with adjustable asymmetric coupling that emulates a Hatano-Nelson model [106].
Fig. 3.
Fig. 3. Topological features of non-Hermitian degeneracies or exceptional points. (a). A topologically protected bulk Fermi arc that connects two exceptional points emerges from a single Dirac point in a 2D-periodic photonic crystal when subjected to radiative losses [148]. (b). Weyl exceptional ring observed in the band structure of a bipartite helical waveguide array with tunable losses [151]. (c) Topological lasing in a coupled waveguide demonstrating dynamic encirclement of an EP [168]. (d) Monitoring the temporal evolution of light pulse during the parametric steering of a non-Hermitian Hamiltonian [169].

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

H = ϵ a n c ^ n a c ^ n a + ϵ b n c ^ n b c ^ n b + n [ κ 1 ( c ^ n b c ^ n a + c ^ n a c ^ n b ) + κ 2 ( c ^ n 1 b c ^ n a + c ^ n a c ^ n 1 b ) ]
H ( t ) = ( i g ~ ( t ) δ ~ ( t ) κ κ i g ~ ( t ) + δ ~ ( t ) )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.