Non-Hermitian physics, which studies systems with nonconservative interactions with the environment, has attracted growing interest. Unique phenomena such as enhanced sensing, the non-Hermitian skin effect, and novel topological behaviors have been extensively explored. Photonic systems serve as ideal platforms to investigate these effects because of inherent decoherence and loss of photons. Non-Hermitian photonics not only reveals intriguing physical effects but also offers new avenues for manipulating light and its associated information, enabling device-level applications. This review summarizes recent progress in non-Hermitian photonic systems. We begin by introducing key concepts and theorems, including the connection between master equations and non-Hermitian Hamiltonians, pseudo-Hermiticity, parity-time symmetry, conserved quantities, exceptional points, and biorthogonal theory. Based on this foundation, we discuss how various platforms—including bulk optics, waveguides, optical cavities, fibers, synthetic dimensions, and metamaterials—simulate non-Hermitian systems. Particular focus is given to the construction of effective non-Hermitian Hamiltonians and operators. We introduce the non-Bloch band theory through a photonic quantum walk platform, highlighting the roles of non-Hermitian topology under complex spectra, which may stimulate advances in both fundamental research and practical applications. Finally, we review unique phenomena and potential applications in sensing, chiral state transfer, quantum algorithms, and other emerging non-Hermitian photonic devices.
【AIGC One Sentence Reading】:Non-Hermitian physics in photonic systems, exploring nonconservative interactions, shows unique phenomena. Photonic platforms are ideal for study. This review summarizes recent progress, covering key concepts, platforms simulating non-Hermitian systems, and potential applications.
【AIGC Short Abstract】:Non-Hermitian physics, exploring systems with nonconservative environmental interactions, has drawn increasing attention. Photonic systems, due to photon decoherence and loss, are ideal for studying unique phenomena like enhanced sensing and the non-Hermitian skin effect. This review summarizes recent advances, covering key concepts such as pseudo-Hermiticity and exceptional points. It discusses how diverse platforms simulate non-Hermitian systems and highlights the non-Bloch band theory, with implications for both research and practical applications.
Note: This section is automatically generated by AI . The website and platform operators shall not be liable for any commercial or legal consequences arising from your use of AI generated content on this website. Please be aware of this.
In quantum mechanics, the observable energy is described by a Hamiltonian operator. According to the postulate of real-valued measurements and the conservation of probability for physical observables, Hermiticity (self-adjointness) is a crucial property of the Hamiltonian when describing a closed system. However, in real physical systems, uncontrolled flows of information, energy, or even particles to the environment lead to a breakdown of probability conservation[1–3]. Generally, the Schrödinger equation with a Hermitian Hamiltonian, which describes the dynamics of a closed system, must be replaced by approaches such as the Lindblad quantum master equation for open systems[4]. Although these methods are extremely powerful, their technical complexity greatly restricts the range of systems that can be efficiently handled. Therefore, the non-Hermitian Hamiltonian, which directly incorporates environmental influences, is proposed as a useful methodology for describing open systems[5–7]. In fact, the non-Hermitian approach and the Lindblad quantum master equation are equivalent in some scenarios. In the short-time limit, typically defined by the inverse of the loss rate[8], the quantum jump terms can be neglected. Effective non-Hermitian Hamiltonians derived from the master equation then provide a good approximation to the dynamics of open systems when post-selection is implemented[9].
Serving as a simple and powerful method for studying open physical systems, non-Hermitian Hamiltonians possess complex spectra, which initially limited their use to purely mathematical tools in earlier research. Driven by curiosity, Bender and Boettcher introduced a broad class of non-Hermitian Hamiltonians exhibiting parity-time () symmetry. These Hamiltonians, characterized by balanced gain and loss, can possess entirely real eigenenergies[10–12]. By abandoning the assumption of Hermiticity, -symmetric non-Hermitian systems have been regarded as a fundamental modification of quantum physics. However, symmetry is not always preserved across all parameter regimes. It can be easily broken by increasing gain and loss, where the eigenenergies transition into complex values. As a result, the properties of the systems differ drastically in the -symmetry-preserving and -symmetry-broken regimes[1,13–17]. Striking examples include the fact that, in the -preserving phases, information can be fully retrieved from the environment, whereas in the -broken phases, retrieval becomes impossible[18–20]. Additionally, under the quench dynamics, quantum dynamic phase transitions accompanied by the emergent momentum-time skyrmions[21] occur in the -preserving phases but vanish in the -broken phases. These significant differences between -preserving and -broken phases have led to intriguing new applications and fascinating phenomena, generating substantial interest over the past decades[22–29]. For instance, in -preserving phases, non-Hermitian systems can enable superluminal information transmission[30]. The exploration of -symmetric non-Hermitian systems has also led to novel applications, such as invisibility[22], non-reciprocal light propagation[23,31], -symmetric laser absorbers[32], and -symmetric phonon lasers[33]. Beyond symmetry, the concept of pseudo-Hermiticity has been proposed to ensure real eigenenergies, with proof that all -symmetric Hamiltonians are indeed pseudo-Hermitian[34,35].
The boundary between -preserving and -broken phases arises at the spectral degeneracies in non-Hermitian systems, known as exceptional points (EPs)[3,35]. In contrast to the conventional energy degeneracy in Hermitian systems, EPs manifest as degeneracies in the spectrum of non-Hermitian Hamiltonian where the corresponding eigenvectors coalesce. This generally makes the system appear as though it suddenly loses the dimensionality of its eigenspace at these critical points[36]. Therefore, while EPs signal -phase transitions, they are also of intrinsic physical interest. A striking feature of EPs is their strong response of the degenerate eigenvalues to perturbations, which makes the system with EPs a good candidate for sensing applications[37–39]. Another immediate consequence of EPs is the general presence of nodal non-Hermitian topological phases, where traditional band-touching points are replaced by these exceptional degeneracies[40,41]. Similar to Weyl semimetals in the Hermitian case, EPs reveal intriguing topologically stable phenomena in non-Hermitian semimetals[42–44]. Furthermore, carrying a nonzero topological charge, EPs exhibit unusual braiding topological and dynamical properties when encircling an EP in the parameter space[45–49]. As one of the unique features of non-Hermitian systems, how to construct higher-order EPs or even exceptional structures is an important topic in this field[50–53]. It has been shown that symmetry is not a necessary condition for the existence of EPs. An th-order EP (a point of n-fold degeneracy in eigenenergies and eigenstates) can be stable existing in a two-dimensional non-Hermitian system without any symmetry. In addition, the required dimensionality for EPs can be reduced through various symmetries. Moreover, these symmetries impart a significantly broader and more conceptually rich range of properties to EPs, which are still being explored[54–56].
Sign up for Photonics Insights TOC Get the latest issue of Advanced Photonics delivered right to you!Sign up now
Besides the aforementioned intriguing aspects of non-Hermitian systems, another fascinating area that has garnered extensive research is the impact of non-Hermiticity on the topological properties of lattice systems. In this domain, symmetries are well known to generally refine the topological classification by constraining the set of permissible physical systems. In Hermitian systems, tenfold symmetry classes are categorized based on the presence or absence of particle-hole symmetry, time-reversal symmetry, and chiral symmetry, as outlined in the comprehensive Altland–Zirnbauer symmetry classification[57]. However, the problem is much more complicated when extending to a non-Hermitian system with the breaking of Hermiticity. In non-Hermitian systems, sublattice and chiral symmetries no longer align. Moreover, non-Hermitian systems possess unique symmetries, such as gain-loss-induced symmetry or pseudo-Hermiticity, which are absent in conventional Hermitian systems. To address this problem, Bernard and LeClair (BLC) developed a 43-fold symmetry classification for non-Hermitian matrices[58]. Building on this foundation, Kawabata et al.[59] proposed a comprehensive theory of symmetry and topology in non-Hermitian physics. Ultimately, the 43-fold symmetry classification is reduced to a 38-fold classification by addressing issues of overcounting and recognizing non-Hermitian symmetry classes.
Topological matter is notable for its remarkable properties and applications, primarily due to the existence of topologically robust edge states. The central principle that relates the bulk topological invariant to these robust edge states is known as the bulk-boundary correspondence (BBC)[60,61]. It says that the number of edge states observed under open boundary conditions (OBCs) corresponds to the bulk topological number defined under the periodic boundary conditions (PBCs). However, a wide range of non-Hermitian systems show markedly different spectra under OBCs and PBCs, suggesting a breakdown of the BBC[62–65]. In addition, in contrast to the Bloch bulk eigenstates in Hermitian systems with both PBC and OBC, eigenstates in non-Hermitian Hamiltonians are highly sensitive to boundary conditions. Although the eigenstates under the PBC remain extended, those under the OBC can all become localized at the boundary, a phenomenon known as the non-Hermitian skin effect[66–68]. To correctly describe this unique phenomenon, the non-Bloch band theory is proposed to rebuild the generalized BBC for non-Hermitian systems[69,70]. The presence of the non-Hermitian skin effect is linked to the existence of a closed loop or a finite region in the periodic-boundary spectrum of one-dimensional or higher-dimensional systems on the complex plane[71–73]. Therefore, the topological origin of the non-Hermitian skin effect is referred to as intrinsic non-Hermitian topology, which can be characterized by spectral windings.
In addition to theoretical advancements, non-Hermitian physics offers fundamental insights for exploring novel phenomena and achieving practical applications across various physical platforms. Non-Hermiticity naturally occurs in a diverse range of systems, including mechanics[51,74,75], acoustics[26,50,76], magnetism[77,78], electrical circuits[79,80], atomic systems[81–83], superconducting qubits[84,85], and photonic systems[14,86,87] (to name a few). In this review, we primarily focus on photonic systems, where non-Hermitian physics has been extensively explored. Without the need for stringent experimental conditions to implement non-Hermiticity, the flexibility of controlling photonic systems makes it an ideal platform for constructing and manipulating non-Hermitian systems. Non-Hermitian physics also introduces a wealth of new functionalities in photonics by expanding the tunable parameter space into the complex domain[88]. Specifically, various photonic platforms, such as the optical waveguides[89], optical cavities[90], and fiber optics[91], are readily available for studying non-Hermitian systems. Integrating non-Hermitian physics into photonics paves the way for new applications, including sensing[37–39], quantum engines[92,93], chiral state transfer[45,46], and state engineering[94], which extend beyond the capabilities of Hermitian photonic systems.
Building on all of these theoretical advancements and optical progress in non-Hermitian physics, we provide this review to give a comprehensive overview of recent developments. The review starts by introducing the basic concepts from a methodological perspective, discussing the relationships between the master equations and non-Hermitian Hamiltonians, -symmetric systems, and conserved quantities. We then review the notable properties of EPs, which are the paramount characteristics unique to non-Hermitian Hamiltonians, to give an intuitive understanding of their generation and application. The concept of biorthogonal eigenstates for the non-Hermitian Hamiltonian will also be addressed in this part. Primarily in this review, we will discuss methods for constructing non-Hermitian systems using different degrees of freedom of photons, such as polarization, spatial modes, time bins, orbital angular momentum, and frequency. Focusing on these advanced methods, we will briefly introduce the implementation of non-Hermiticity in photonic platforms of bulk optics, optical waveguides, optical cavities, fiber optics, synthetic dimension in photonics, and optical metamaterials. Our review further covers the progress on the applications of non-Bloch band theory in demonstrating the unique phenomena of skin effects and edge bursts in non-Hermitian systems. In addition, we will outline the key concepts of non-Hermitian linear response theory, non-Bloch EPs, and the reconstruction of the generalized BBC with the generalized Brillouin zone to illustrate the power of the non-Bloch band theory. We will also review emerging non-Hermitian topological matters and their simulations in photonics. The novel functional applications induced by non-Hermitian physics in photonics will also be presented in this review. Finally, the review will conclude with a summary and outlook on the future of non-Hermitian physics in photonic systems.
2 Fundamental Theory and Background of Non-Hermitian Physics
2.1 Master Equations
In the framework of classical quantum mechanics, the temporal evolution of an isolated system is fundamentally described by the Schrödinger equation where represents the state of the system at time , and denotes the Hamiltonian of the complete system, which acts on the corresponding Hilbert space.
A fundamental postulate of quantum theory states that the Hamiltonian of an isolated system must be Hermitian. This Hermiticity guarantees real, single-valued energy eigenvalues and ensures a unitary time evolution of the system. However, the real physical system is called an open system since it is interacting with an environment so that the full dynamics of the system and environment are unitary[95,96]. Thus, for an open system interacting with an environment, the description of physical states through wave functions as in Eq. (1) is no longer complete when dealing with composite systems[97]. Here, we use the density operators, which are in particular an appropriate description of subsystems to describe physical states instead of wave functions in general[98]. The Schrödinger equation of an open system turns into where is the corresponding density operator of composite systems that contain a real physical system and the environment . The Hamiltonian , with representing the Hamiltonian of the interaction between two subsystems[99]. Assuming there is no interaction between and at , the density matrix can be decomposed into [100]. The density operators representing the state of the subsystems () are obtained by tracing out (), respectively. Therefore, the reduced dynamical equation for the physical system can be written as
In the interaction picture, the evolution equation of the composite system is , with and . Iteratively substituting the system’s density operator into the evolution equation and taking the partial trace over the environment on both sides of the equation, the following formula can be derived:
The expected value of the coupling observable is zero under the initial state since there is no interaction between and at , which ensures that . As time goes on, the interaction between the physical system and the environment involves a connection. Assuming the interaction is weak, that is, under the weak-coupling approximation or the so-called Born approximation[95], the environment is virtually unaffected by the physical system, which gives rise to the fact that the environment remains in its initial state at any time. Then, the density operator of composite systems in Eq. (4) can be approximated as .
The Hamiltonian of the interaction between two subsystems can be analyzed using a low-order expansion over time . However, the specific form of the density operator remains highly complex, even though the evolution of the physical system is non-Markovian indeed. Fortunately, if the rate of environmental re-equilibration is very fast, the time scale over which the state of the system varies appreciably is large compared to the time over which the environment correlation functions decay. In this case, the environment is approximately equivalent to erasing the memory of the system, which leads to the results that the state of the physical system can be regarded as determined by the current time, that is, . This method is called Markov approximation[95]. Last, assuming the environment is composed of a series of resonators and , with representing the boson operator, the interaction Hamiltonian in the interaction picture can be written as , where is the creation operator. By introducing the rotating-wave approximation for the interaction and switching back to the Schrödinger picture, the master equation for the physical system can be written as[95]where represents the transition, dissipation, or decoherence rate of the corresponding interaction channel. is called jump operators, which represent the coupling between the physical system and the environment.
Equation (5) is called the Lindblad master equation and it describes the dynamics of an open quantum system interacting with an environment[4,101]. is the generalized Liouvillian operator, which can be regarded as the generalization of the Hamiltonian in open systems. The first term in Eq. (5) represents the unitary evolution, and the other terms describe the changes that occur in the physical system as a result of the interaction between the physical system and the environment.
2.2 Non-Hermitian Hamiltonian
From the Lindblad master equation, it is clear that the dynamics of the quantum states are not only governed by the Hamiltonian of the physical system, but by additional interaction terms posed by the environment. Although a large variety of physical systems are well described by this equation, such as the atomic systems and quantum optical systems, the solutions to the master equation are often complicated. Here, we introduce a unique way to address the problems with various kinds of open systems. By constructing a complex quantum potential, the Lindblad master equation in Eq. (5) can now be rewritten as
The second term in Eq. (6) represents the quantum jumps, which can be considered as decoherence arising from the interaction between the physical system and the environment. The remaining gain-loss terms are incorporated into the system’s evolution, making it non-Hermitian.
The occurrence of quantum jumps requires a time accumulation process; thus, their effects on the system may not be significant within extremely short time windows. Additionally, if the system resides in a low excited state or if the dissipation rate is much smaller than the system’s coherent evolution rate, neglecting quantum jumps becomes reasonable. For example, in cavity quantum electrodynamics systems, when the cavity’s loss rate is low and the measurement duration is extremely short, a non-Hermitian Hamiltonian can be used to approximate the cavity mode decay[102]. In this case, photon escape is treated as continuous loss, and ignoring quantum jumps may lead to theoretical predictions aligning well with experimental observations on short timescales.
However, over longer time periods or when the system is in a highly excited state, the cumulative effects of quantum jumps become significant, and their occurrence frequency increases. Ignoring jumps in these scenarios can result in inaccurate predictions. Furthermore, in photonic experiments, non-Hermitian Hamiltonians induce effective losses or gains, which may alter the system’s spectrum. Ignoring quantum jumps may affect photon pair generation efficiency or introduce discrepancies in photon statistics measurements. For instance, in photon pair correlation measurements, non-Hermitian models may predict different second-order coherence functions, but actual quantum jumps introduce additional noise or decoherence effects[103]. While the evolution of the average photon number may align with the predictions of a non-Hermitian Hamiltonian, higher-order correlation functions may deviate significantly. Moreover, in entanglement measurements, neglecting quantum jumps may overestimate the entanglement lifetime, as actual dissipation destroys entanglement, whereas the non-Hermitian model may only account for amplitude decay, ignoring the collapse of the state due to quantum jumps.
Here, we analyze the cases where the effects of quantum jumps can be neglected. By ignoring the quantum jumps of the composite systems, the density matrix of the physical system can be considered to approximate a non-Hermitian Hamiltonian[95]:
One typical feature of the non-Hermitian system is that it cannot have both real energy spectra and mutually orthogonal eigenvectors, that is, the energy of the physical system can be complex. The complex energy corresponds to the time evolution . It can be clearly seen that the imaginary part of the energy corresponds to a gain or loss. Thus, the non-Hermitian system can be realized by applying controlled gain-loss to the Hermitian system in general[104,105].
2.3 Symmetry
While generic non-Hermitian systems are characterized by complex energy spectra and the non-orthogonality of eigenstates, further constraints on the Hamiltonian structure can lead to remarkable spectral properties. Among them, symmetry—where the combined action of parity () and time-reversal () commutes with the Hamiltonian—has emerged as a particularly fruitful framework. Although there have been some studies on real energy spectra of non-Hermitian Hamiltonians in different fields at an early stage[106–112], a complete theory has not yet been formed. Until 1998, Bender and Boettcher proposed a special series of -symmetric Hamiltonians and elaborated the relationship between symmetry and real energy spectrum[10]. As a result, non-Hermitian quantum mechanics began to gain widespread attention. In 2002, Mostafazadeh constructed the -pseudo-Hermitian symmetric quantum system based on a biorthogonal basis mathematically, and explained the conditions required for the real energy spectrum of non-Hermitian Hamiltonians[113–115]. Since then, generally speaking, non-Hermitian quantum mechanics has been largely discussed along the two branches of -symmetric and -pseudo-Hermitian quantum systems, as we discuss in the following.
The discussion of -symmetric systems was initiated by a deformed harmonic-oscillator Hamiltonian[10,116,117]: where is the momentum operator and is the position operator. This Hamiltonian is a complex non-Hermitian but has a real energy spectrum while most people used to think that only Hermitian operators had this property previously. It is further found that in Eq. (8) is symmetric, that is, satisfies the following equation: where and are the parity operator and time reversal operator, respectively, and . The parity operator is a linear Hermitian operator that performs a space reflection and , , so under parity reflection both and change sign: , . The time-reversal operator is an anti-linear reflection operator and for bosonic systems , . Under time-reversal reflection, it can be shown that , .
In order to reveal whether the real energy spectrum of the non-Hermitian Hamiltonian is related to symmetry, Bender applied a special perturbation method to expand the -symmetric Hamiltonian in terms of , and replaced with [10,116]. Here, is a small real number and does not change the symmetry of the non-Hermitian Hamiltonian above. Therefore, focusing on the formula of the replaced Hamiltonian, the range of is further extended and processed by numerical calculation. Finally, according to the different values of , he obtained the solution of the system with real and conjugate complex eigenvalues.
The publication of Ref. [10] began to arouse people’s attention to non-Hermitian quantum theory, and studies based on the real energy spectra of -symmetric Hamiltonians are appearing with increasing frequency[116,118–125]. In addition, the symmetry is studied theoretically from the mathematical point of view[126–133]. Among the many novel characteristics of non-Hermitian -symmetric systems, the most noteworthy is the case that the -symmetric Hamiltonian has real energy spectra. Since in traditional quantum mechanics, a Hermitian Hamiltonian with positive real energy spectra corresponds to an observable, a non-Hermitian Hamiltonian with real energy spectra can also theoretically be regarded as an observable. In order to give a self-consistent interpretation to such a non-Hermitian system in physics, several conditions must be satisfied: the positive definiteness of the inner product of the system, the completeness of the system, and the unitarity of the time evolution.
The eigenstates of -symmetric non-Hermitian Hamiltonians do not satisfy the standard completeness relations[134–137]. We can define as energy eigenstates of a particle on the real line subjected to a -symmetric potential. Unlike traditional quantum mechanics where we know what the inner product is even before the given Hamiltonian, for a -symmetric Hamiltonian, a plausible guess for what the inner product might be should be considered first. Suppose the energy eigenstates associated with -symmetric systems are also orthogonal; numerical evidence implies that the inner product has been proposed[138,139]: Here we replace the symbol with in order to distinguish the Hermitian Hamiltonian. The coefficient originates from the nonstandard completeness relations of the -symmetric non-Hermitian Hamiltonian , which is mathematically proven in Ref. [138]. The function is the energy eigenstate of a particle on the real line subjected to a -symmetric potential and is the eigenvalue of the th energy level. The negative norm makes the metric of the inner product of -symmetric systems uncertain, and it is difficult to maintain the familiar probabilistic interpretation of quantum theory. However, the inner product has the same number of positive and negative terms that are alternating; thus, the extra implies that there should be some unknown hidden symmetry within the system.
In an attempt to construct an inner product endowed with a positive-definite norm and establish that the -symmetric Hamiltonian defines a unitary quantum mechanical theory, a remedy against the indefinite metric in Hilbert space has been proposed in the form of a linear charge operator , whose properties closely resemble those of the charge conjugation operator in particle physics[140], known as the theorem.
In coordinate space, the linear operator , which represents the additional symmetry of -symmetric Hamiltonian, is a sum over the -normalized eigenfunctions and satisfies[141]
It is easy to verify that and . Therefore, the operator is a measure of the inner product for the -symmetric Hamiltonian, and satisfies . Thus, by introducing into the inner product, the negative signs therein can be cancelled out. The advantage of the inner product is that, unlike the inner product, the inner product is positive definite, and the completeness condition of the -symmetric system in coordinate space is rewritten as[138]
Last, the coherent, non-unitary time evolution operator for the system based on the Schrödinger equation is given by
The time evolution operator is the same as the Hermitian quantum system, except that the Hermitian Hamiltonian is replaced by the -symmetric Hamiltonian. Consider as a given initial state; the final state at is defined as . It is straightforward to calculate that the norm of the final state Here, is the matrix transpose and . It can be seen that the unitarity of the time evolution is also guaranteed in -symmetric systems.
In traditional Hermitian quantum mechanics, a linear operator must be Hermitian to be an observable, that is, . This condition is obtained by complex conjugation in positive definite inner products of Hermitian operators, which ensures that the expectation value of in a state is real. Similar to the Hermitian operators, in -symmetric quantum mechanics the condition for an operator to be an observable is[141]
This condition also guarantees that the expectation value of in any state is real since . It is easy to prove that the operator and the -symmetric Hamiltonian themselves satisfy Eq. (15), so they are all observables.
Besides, the operator satisfies , which is an extremely strong constraint. Otherwise, the norm of the system would not be guaranteed to be conserved under a set of orthogonal positive definite inner products. However, there exists in other non-Hermitian systems. For this case, the pseudo-Hermitian inner product or biorthogonal basis inner product needs to be introduced. For example, in Ref. [142], Jones obtained the pseudo-Hermitian inner product of the system by transforming the wave function of the harmonic oscillator when he analyzed the Swanson Hamiltonian with symmetry but transposed asymmetrically. We will discuss the biorthogonal quantum mechanics and pseudo-Hermitian systems in the next subsection in detail.
2.4 Biorthogonal Quantum Mechanics and Pseudo-Hermitian Systems
-symmetric systems show that non-Hermitian Hamiltonians can still have real spectra under certain conditions. However, due to the non-Hermiticity, their eigenstates are generally not orthogonal, and the usual rules of quantum mechanics no longer apply directly. To handle these issues, a more general framework is needed. This leads to biorthogonal quantum mechanics and pseudo-Hermitian systems, which provide consistent ways to describe measurements, inner products, and time evolution in non-Hermitian settings.
Compared to the Hermitian systems with real eigenvalues and orthogonal eigenstates, the eigenvalues and eigenstates in a non-Hermitian Hamiltonian are not necessarily real and orthogonal. In biorthogonal quantum mechanics[131,143–146], the orthogonality of eigenstates is replaced by the notion of biorthogonality that defines the relation between the Hilbert space of states and its dual space. Consider a non-Hermitian Hamiltonian ; the eigenvalue equations of and with the th energy level are given by where and are the th eigenvalues, and and are the right eigenstates of and , respectively. Using the transpose complex conjugation on both sides of Eq. (16), it can be obtained that
Since the energy spectra are unique, it is straightforward to obtain the biorthonormal relation[146]Here and are two different sets of eigenstates that correspond to and due to . Equation (18) extends the orthogonality in traditional quantum mechanics. Accordingly, the completeness relation of non-Hermitian systems should also be generalized. An arbitrary quantum state can be expressed by the eigenstates as
Thus, the probability amplitude is calculated as using the biorthonormal relation. Obviously, the state can be rewritten as
Therefore, the right and left eigenstates satisfy the completeness relation[146]
In summary, the aforementioned relations are two basic properties of biorthogonal quantum mechanics in non-Hermitian systems, and are the so-called biorthogonal eigenstates.
The eigenstates of a biorthogonal system are no longer normalized under the completeness relation, which leads to the fact that the traditional probabilistic interpretation used in Hermitian quantum mechanics has changed. Because it is straightforward to calculate that , by assuming that all eigenstates have the same Hermitian norm for all . In this case, the inner product should be redefined by introducing the so-called associated state that defines duality relations between elements of the Hilbert space and its dual space.
For an arbitrary state in Eq. (19), the associated state is defined according to the following relation: where the state associated with is then given by the Hermitian conjugate of the bra vector with . Thus, by expressing another state with the same eigenvectors, that is, , the inner product for a biorthogonal system is defined as[146]
Next, we will discuss the pseudo-Hermitian Hamiltonian system based on the biorthogonal quantum mechanics[113–115]. A Hamiltonian is pseudo-Hermitian with respect to if it satisfies where is called an intertwining operator, which is generally a linear Hermitian invertible transformation[147].
In order to observe the equivalence of the existence of a complete biorthonormal set of eigenstates of , we introduce the degeneracy label into the completeness relation and the biorthonormality relation. Naturally, the evolution of the state can be written as[113–115]
Thus, it can be seen that is the eigenstate of corresponding to the eigenvalue . Moreover, the operator maps the eigensubspace with eigenvalue to the subspace with eigenvalue . The eigenfunctions corresponding to these two eigenvalues have the same degeneracy.
Furthermore, using the completeness relation in Eq. (21), it is easy to obtain the transition relation between the two biorthonormal eigenstates and as follows:
Therefore, the complete biorthonormal set of eigenstates of pseudo-Hermitian Hamiltonian based on the indefinite inner product is given by[113]
Besides, it can be easily calculated to find the form of the invertible linear operator[113]: Here, is the degree of degeneracy of eigenvalue . Meanwhile, the pseudo-Hermitian Hamiltonian can be expressed by
It is clearly seen that the expansion based on a biorthogonal basis is self-consistent with the pseudo-Hermitian theory. The above analysis leads to two properties: i) is a pseudo-Hermitian Hamiltonian if and only if the eigenvalues are real or come in complex conjugate pairs; ii) if is pseudo-Hermitian with respect to two linear Hermitian invertible transformations and , then the intertwining operators satisfy .
Now consider a -symmetric Hamiltonian that has the discrete energy spectrum with a set of biorthogonal bases. After using the eigenvalue equation , the commutation relation is used. The formula is changed into
Obviously, is the eigenvalue of corresponding to the eigenstate . The energy spectra are real in the -symmetric region and come in complex conjugate pairs in the -symmetry broken region. Thus, a -symmetric Hamiltonian system with discrete energy spectra and a set of biorthogonal bases belongs to the pseudo-Hermitian category.
The above pseudo-Hermitian theory is mainly based on the biorthogonal states with discrete energy spectra. However, people are more accustomed to discussing pseudo-Hermitian theory in the framework of a set of orthogonal bases. Consider to be a set of orthogonal bases in Hilbert space; the indefinite inner product of two arbitrary bases is defined as[148]
The indefinite inner product is an extension of the ordinary Hermitian inner product. It degenerates into the Hermitian inner product when ; meanwhile, the pseudo-Hermitian Hamiltonian changes into Hermitian.
The evolution of the pseudo-Hermitian system with initial state is governed by the Schrödinger equation
It is easy to check that the equation for the indefinite inner product is given by
Thus, it can be concluded that the probability of indefinite inner product space defined by a pseudo-Hermitian Hamiltonian is a conserved quantity.
Similar to the calculation above, based on the eigenvalue equation, the eigenvalues of the pseudo-Hermitian Hamiltonian satisfy
Therefore, the pseudo-Hermitian systems based on indefinite inner products can only have real energy spectra. Otherwise, when , the constraint condition of the above equation becomes , which means that the inner product of the same state is zero under any circumstances, thereby violating the probability explanation in quantum theory obviously.
2.5 Exceptional Points
In non-Hermitian systems, the energy spectra can be complex with the corresponding eigenstates being non-orthogonal and the corresponding time evolution operators are non-unitary. When the skewed eigenstates completely coalesce at certain points in the parameter space, the eigenenergies also become degenerate[149]. In this case, the non-Hermitian Hamiltonian cannot be diagonalized and the eigenspace collapses to a lower dimension and becomes incomplete, which is analogous to the reduction of the dimension of Hilbert space[3,36]. The special points are referred to as the EPs and the number of degeneracies is the order of EPs.
It is often accompanied by novel phenomena such as ultrahigh sensitivity[37,38] and phase transitions[150,151] in the immediate vicinity of an EP. The special algebraic behavior allows a reduction of the full problem to the two-dimensional problem associated with the two coinciding levels[152,153]. Here consider the two-dimensional Hamiltonian
The Hamiltonian becomes non-Hermitian when either parameter or is subjected to analytic continuation. In an open system, for simplicity, we can set and , where may be realized as detuned energy, and denotes the loss (gain) of the system. It is straightforward to calculate the eigenvalues and the eigenstates as
The eigenvalues and the eigenstates are identical at the EP, where , and the Hamiltonian becomes a defective matrix. Interestingly, the eigenstates at the EP can be mapped to the Jones vector describing circular polarization of light, which is referred to as chirality[154].
Note that at the EP, the difference between degeneracy and coalescence is clearly manifested by the occurrence of only one eigenstate instead of the familiar two in the case of a genuine degeneracy. Using the biorthogonal system for a non-Hermitian Hamiltonian in the previous subsection, the inner product of the eigenstates at EP satisfies , which means that the eigenstates at EP are self-orthogonal[155]. It is the vanishing of the inner product that enables the reduction of a high-dimensional problem to two dimensions in the close vicinity of an EP[154].
In fact, even though the Hamiltonian at EPs is not diagonalizable, it can be brought to Jordan normal form via a similarity transformation , with being the block diagonal, consisting of Jordan blocks[156,157]. Then, the existence of higher-order EPs depends on checking whether the eigenvalues and their corresponding eigenstates of the Jordan blocks coalesce. Note that EPs of different orders can coexist since the Hamiltonian may feature multiple such Jordan blocks of varying dimensions.
In the non-Hermitian systems without symmetry, finding an th-order EP requires tuning parameters, meaning that the appearance of higher-order EPs requires increasing the number of fine-tuned parameters, and they are thus assumed to play a much less prominent role[158]. However, Mandal and Bergholtz proposed that third-order EPs generically require only two real tuning parameters in the presence of either a symmetry or a generalized chiral symmetry[54]. This shows that the requirement for tuning parameters to find an th-order EP can be reduced in non-Hermitian systems by introducing different symmetries. However, the constraints required for higher-order EPs under different symmetries remain an open problem.
Recent interest in higher-order EPs, where more than two eigenstates coalesce, has motivated significant efforts in their realization and application. While second-order EPs are relatively well-understood and widely demonstrated, constructing third- or higher-order EPs typically requires the fine-tuning of multiple parameters to enforce the degeneracy condition. A general th-order EP satisfies the condition , where all eigenvalues and eigenvectors coalesce. This gives rise to a characteristic eigenvalue sensitivity scaling as under perturbation strength , offering potential advantages for sensing. In photonic systems, such EPs can be realized by coupling multiple microring resonators[159,160], photonic crystal defect cavities[161], or by exploiting synthetic dimensions, such as dynamically modulated or time-multiplexed structures[162]. These approaches effectively construct higher-dimensional non-Hermitian Hamiltonians, enabling the coalescence of multiple modes.
Importantly, certain symmetries can significantly simplify the realization of higher-order EPs by reducing the number of tuning parameters. In particular, symmetry allows real spectra to persist despite non-Hermiticity, and facilitates EP formation under constrained conditions. For example, in a three-mode -symmetric system composed of gain, loss, and neutral components, only two independent parameters may suffice to reach a third-order EP[163,164]. Likewise, chiral (or sublattice) symmetry enforces a symmetric spectrum about zero, leading to a block-off-diagonal Hamiltonian structure of the form . This naturally supports pairwise eigenvalue coalescence, aiding in the emergence of higher-order EPs with reduced tuning effort[165]. These symmetry-induced constraints lower the codimension of the EP manifold, enhancing both the feasibility and stability of experimental realizations.
Despite these theoretical insights, observing the enhanced sensitivity associated with higher-order EPs remains experimentally challenging. While the scaling implies stronger response to target perturbations, it also amplifies the effects of unwanted noise, disorder, and fabrication imperfections[166]. Moreover, the coalescence of eigenvectors at an EP leads to non-orthogonality, which can degrade the signal-to-noise ratio (SNR) and limit precision in practical measurements[167]. Recent theoretical analyses further show that the quantum Fisher information (QFI), which bounds estimation precision, may saturate or even decline near EPs under realistic decoherence, thus challenging the assumption that EPs always enhance metrological performance[168,169]. These issues highlight the need for noise-resilient architectures, dynamically tunable platforms, and hybrid schemes that can robustly harness the non-Hermitian features of higher-order EPs without sacrificing precision.
2.6 Conserved Quantities
The conserved quantities of a closed system are determined by its symmetries and offer global insights into its dynamical evolution, such as energy, the Runge–Lenz vector in Kepler dynamics[170], or the electric charge[171]. Constraints imposed by these time invariants have been instrumental in advances ranging from the prediction of neutrinos to the development of conserving approximations in many-body physics.
In closed systems, an observable is called conserved if it commutes with the Hermitian Hamiltonian of the system. Due to the equivalence between conservation and commutation, a unitary symmetry transformation on the quantum state space is generated by each conserved observable . One can construct simultaneous eigenstates of the Hamiltonian and the observable such that the Hamiltonian becomes block-diagonalized, thereby reducing the complexity of the eigenvalue problem. A complete set of conservation laws for a system is obtained by identifying all linearly independent observables that commute with the Hamiltonian.
However, for a non-Hermitian Hamiltonian , the norm of the wave function is not conserved. A complete characterization and observation of conserved quantities in open systems and their consequences is still an outstanding question. With an approach inspired by the pseudo-Hermiticity[113–115], these questions can be addressed. If is an intertwining operator, i.e., it satisfies the relation , it is straightforward to show that is constant for any arbitrary states and at any arbitrary time . When the Hamiltonian is Hermitian, the intertwining relation has a trivial solution . This implies that the Dirac inner products, and in particular, Dirac norm, are time invariant. All other nontrivial solutions correspond to operators that commute with the Hamiltonian and thus correspond to conserved quantities in the traditional Hermitian quantum theory.
When , the non-unitary time evolution operator keeps the observable unchanged, that is, . Thus, an expectation value of in an arbitrary quantum state is a conserved quantity. The equivalence between conservation and commutation breaks down for a non-Hermitian Hamiltonian, meaning that the conserved observables and the Hamiltonian cannot be simultaneously diagonalized[172].
The defining equation for the intertwining operator in open systems allows one to systematically solve for it. Note that since this process is linear, if is a solution and is another solution, then their linear combinations are also intertwining operators. Thus, one needs a suitable definition of to satisfy the conditions of orthogonality and linear independence.
Consider a -symmetric Hamiltonian for instance; the form of is given by where and are the spin operators that satisfy the canonical commutation relations . The simplest representation for the generators of SU(2) is the spin representation. This is given by two-dimensional matrices , where are the three Pauli matrices. The analytical recursive construction is as follows[172]. The transpose symmetry of the -symmetric Hamiltonian implies , where the time-reversal operator is given by complex conjugation . It follows from the symmetry of the Hamiltonian that the parity operator is a conserved observable, namely, . Therefore, a sequence of linearly independent and dimensionless observables can be constructed by . The intertwining nature of implies that is also a conserved observable. This sequence terminates with since the characteristic equation for is a polynomial of dimensional order [147].
For , it is easy to check that are two linearly independent but not orthogonal intertwining operators. Thus, their expectation values are independent of time, irrespective of whether the system is in the -symmetric region or the -broken region. Furthermore, a coherent superposition of eigenstates of a conserved quantity cannot be generated from a single initial eigenstate in an isolated system. However, the eigenstates of a conserved observable exhibit nontrivial dynamics in -symmetric systems[172].
3 Experimental Implementation of Non-Hermitian Physics in Photonic Systems
Although non-Hermitian physics originates from quantum mechanics, optical systems have emerged as an ideal platform for its study. This is primarily due to the ability to control gain and loss effects in open systems by manipulating components within the optical system. The characteristics of non-unitary evolution are investigated by realizing non-Hermitian Hamiltonians. Non-Hermitian systems give rise to novel physical phenomena due to their diverse symmetries and exceptional points, offering new solutions for information processing, topological structures, precision measurement, and other areas of research.
Non-Hermitian physics has already been experimentally realized in various optical systems. It is helpful to briefly reflect on the distinctive advantages that non-Hermitian physics brings to photonic systems. Beyond theoretical interest, non-Hermitian effects offer new degrees of freedom—such as loss engineering, complex-valued coupling, and asymmetric mode interactions—that have led to a range of novel functionalities in photonics. For example, by leveraging asymmetric coupling or gain-loss control, non-Hermitian photonics has enabled the realization of mode-tunable lasers, where the lasing mode can be selectively addressed through parameter tuning near exceptional points[173,174]. Similarly, sharp phase transitions and nonreciprocal dynamics in non-Hermitian systems have been harnessed for ultrafast and compact optical switches[175]. Furthermore, by embedding non-Hermitian terms in topologically structured lattices, researchers have demonstrated reconfigurable topological photonic systems, allowing for flexible control of edge transport and robust light steering[176,177].
These examples underscore the practical relevance of non-Hermitian models in photonics and motivate their continued experimental exploration. In this section, we examine the application potential of non-Hermitian physics in optical platforms by discussing its realization in different optical systems, including bulk systems, time-multiplexed systems, waveguides, optical lattices, optical cavities, optical metastructures, and optical combs.
3.1 Implementation of Non-Hermitian Physics in Bulk Optics
Non-Hermitian Hamiltonians with symmetry exhibit real energy spectra, which have garnered significant attention in classical optical systems[178–187] and have been employed in various applications[188–193]. The currently reported realization of -symmetric synthetic optical lattices with balanced gain and loss relies on classical light pulse propagation, analogous to quantum walks[194].
Quantum walks[195–199], as a versatile quantum simulation framework, describe the movement of particles with internal degrees of freedom on discrete lattices. However, implementing this approach in the quantum domain poses challenges, particularly due to the difficulty in amplifying single photons. In 2017, Xiao et al. experimentally realized passive -symmetric single-photon quantum walks[200], where photon losses were modulated intermittently in a quantum walk interferometer network. By alternately increasing and decreasing these losses, the study observed the dynamical behavior of the non-Hermitian system in regimes where symmetry was both preserved and broken. They realize passive -symmetric non-unitary quantum walks with single photons by employing a scheme that alternates photon losses[201–204]. The dynamics of the quantum walk are dictated by a non-unitary time-evolution operator , where is the conditional translation operator, and is the position-dependent coin operator. The symbol is the position of walker and are the orthogonal coin states. is the loss operator.
As illustrated in Fig. 1, the spatial modes of photons encode the walker states, while their horizontal and vertical polarization states define the coin states. A position-dependent coin operator is achieved using individually placed unmounted half-wave plates (HWPs) along designated spatial paths. Conditional position shifts are introduced by birefringent beam displacers, which transmit vertically polarized photons straight through while displacing horizontally polarized photons laterally to an adjacent mode. The polarization-dependent loss operator is implemented using a single-input partially polarizing beam splitter (PPBS) with distinct transmission coefficients for horizontal and vertical polarizations. The raw probabilities are calculated by normalizing the coincidence counts against the total number of photon pairs generated prior to interacting with the experimental setup at each step.
Figure 1.Experimental setup for non-unitary quantum walks with alternating losses[200]. A photon pair is generated in a Bell state through spontaneous parametric down-conversion. One photon serves as a trigger and can be prepared in any desired polarization state using a quarter-wave plate (QWP), a half-wave plate (HWP), and a polarizer for projection measurement. The other photon undergoes a series of processes, including controlled polarization rotations, polarization-dependent loss operations, and spatial translations. To explore the topological characteristics of the quantum walk, unmounted HWPs are strategically placed along specific paths to introduce spatially varying polarization rotations. The resulting probability distribution is recorded using avalanche photodiodes.
3.2 Implementation of Non-Hermitian Physics in Time-Multiplexed Systems
In the realization of quantum walks, time encoding of photon arrival times at the detector can be employed in addition to encoding in spatial degrees of freedom[205–209]. This approach overcomes the limitations of physical space resources, enabling more steps to be implemented and facilitating the observation of complex quantum dynamics. Time encoding eliminates the need for additional spatial resolution devices or a large number of optical components, such as beam splitters and delay lines, resulting in a more compact experimental setup with reduced complexity and minimized error accumulation. Furthermore, time encoding is more readily extendable to high-dimensional quantum walks. For instance, by introducing multiple cascaded time delays, multi-dimensional quantum walks can be achieved without being constrained by the size of the experimental platform. An additional advantage is the reduced sensitivity to noise. Since the position of a single detector remains fixed and does not require complex optical alignment, environmental interference has a minimal impact on the signal, improving the signal-to-noise ratio of the experiment.
The experimental realization of non-Hermitian quasicrystals via quantum walks employs a sophisticated photonic setup that leverages single-photon dynamics and time-multiplexed techniques[207]. As shown in Fig. 2, the experiment begins with a pulsed laser emitting photons at a central wavelength of 808 nm, a pulse width of 88 ps, and a repetition rate of 31.25 kHz. The laser pulses are attenuated to the single-photon level using a neutral density filter, ensuring an average photon number per pulse of less than . This guarantees a negligible probability of multi-photon events, which could otherwise introduce unwanted noise. The photons are then directed through a polarizing beam splitter (PBS) and an HWP to initialize their polarization states. This step encodes the coin degrees of freedom of the quantum walk into the photon’s polarization, providing a robust platform for simulating discrete-time quantum walks.
Figure 2.(a) Conceptual illustration of the quantum walk dynamics[207]. The shift operator and coin operator are depicted, with lattice sites represented at the intersection points of the laser beams (red). Quasiperiodic phase modulation is introduced by the phase operator , illustrated as the dashed gray curves. (b) Detailed experimental setup using a time-multiplexed architecture. Key components include: a pulsed laser source, a neutral density (ND) filter for attenuation, HWPs for polarization control, an electro-optic modulator (EOM) for phase modulation, polarizing beam splitters (PBSs) for routing photons, and avalanche photodiodes (APDs) for detection. The two fiber loops with different lengths encode the walker’s position in the time domain, enabling the simulation of non-Hermitian quasicrystals.
The quantum walk dynamics are implemented using a time-multiplexed setup, where the walker’s position is encoded in the time domain. The setup consists of two optical fiber loops of slightly different lengths (167.034 and 160.000 m), coupled via PBSs. The shift operator is realized by separating photons into their horizontal and vertical polarization components and routing them through the two fiber loops. The temporal separation between the photons traveling through the loops (33.046 ns) defines the time-bin width, effectively mapping the walker’s position onto discrete time intervals. The coin operator , which governs the polarization state transitions, is implemented using two HWPs that precisely control the parameter . The position-dependent phase operator is realized using an electro-optic modulator (EOM), which introduces a phase shift . The EOM’s fast rise/fall time (4 ns) ensures accurate phase modulation within the temporal resolution of the setup. Together, the coin and phase operators enable the simulation of a non-Hermitian quasicrystal with tunable quasiperiodic potentials and non-reciprocal hopping.
To implement a polarization-dependent loss operation , two HWPs are introduced into each fiber loop. For the shorter loop, horizontally polarized photons are transmitted by the PBS and continue their evolution, while for the longer loop, a fraction of vertically polarized photons are flipped into horizontal polarization and subsequently lost from the system. This controlled loss mechanism, governed by the parameter , is crucial for simulating non-Hermitian dynamics in the quantum walk. Photons exiting the system are detected using three avalanche photodiodes (APDs), which measure the temporal and polarization properties of the out-coupled photons. By analyzing the photon counts at each time step, the probability distribution of the walker is reconstructed. This allows for the calculation of key observables, such as the dynamic inverse participation ratio (dIPR), which provides insights into the localization properties of the system.
3.3 Implementation of Non-Hermitian Physics in Waveguide Systems
In 2013, Crespi et al.[210] demonstrated quantum walks using integrated interferometer arrays fabricated in glass through femtosecond laser writing. This technique enabled the introduction of controlled phase shifts at each unit cell of the network[13,189,211–216]. The approach holds significant potential for quantum simulation and for achieving computational capabilities that surpass those of conventional computers in “hard-to-simulate” scenarios. The authors also investigated the interaction between the Anderson localization mechanism and the Bosonic/Fermionic symmetry of the wave function in waveguide circuits.
Femtosecond laser waveguide writing, as illustrated in Fig. 3[210], employs the nonlinear absorption of femtosecond pulses focused just beneath the surface of a transparent dielectric substrate to achieve a permanent local increase in refractive index. This method allows for the precise fabrication of optical waveguide circuits along a desired path, enabling arbitrary three-dimensional designs. It offers significant advantages in speed and versatility compared to traditional photolithography, eliminating the need for photomasks. In the system depicted in Fig. 3(a), the discrete -axis represents the quantum walk sites and the -axis corresponds to time steps. Phase shifts are controlled by deforming one of the S-bent waveguides at each directional coupler output, stretching the optical path. The phase shift range is achieved by modifying the optical path length in either the green or red segments, corresponding to phase shifts of and , respectively. This approach allows small, consistent deformations to cover the full range of phase shifts. Figure 3(c) compares undeformed and deformed S-bends. To calibrate the phase shift as a function of deformation, several Mach–Zehnder interferometers, as shown in Fig. 3(d), were fabricated, each with one S-bend deformed by a different value of the deformation parameter . The experimental results of phase shifts as a function of are presented in Fig. 3(e).
Figure 3.Integrated circuit for disordered quantum walks[210]. (a) A schematic of a network of directional couplers is shown, implementing an eight-step one-dimensional quantum walk with static disorder. Different phase shifts are represented by distinct colors, while violet waveguides indicate accessible paths. (b) The optical path length is controlled by deforming one of the two S-bent waveguides at the output of each directional coupler, effectively functioning as a phase shifter. (c) The deformation is modeled by a nonlinear coordinate transformation dependent on a deformation coefficient. The plot compares the undeformed (solid line) and deformed (dashed line) S-bends. (d) A Mach–Zehnder structure is used as the unit cell for the directional coupler network, which is fabricated for calibrating the phase shift induced by the deformation. (e) The phase shift induced by the deformation is presented, showing the theoretical curve (solid line) calculated from the geometric deformation and experimental data (diamonds) confirming the results.
3.4 Implementation of Non-Hermitian Physics in Photonic Lattices
The construction of disordered photonic lattices is a cornerstone for studying Anderson localization and other wave phenomena in optics. In 2007[217], a robust and versatile method was employed to generate a two-dimensional periodic photonic lattice with random perturbations, leveraging advanced optical techniques and precision control[218–222].
As shown in Fig. 4, the periodic lattice was induced in a photorefractive crystal (SBN:60) using a structured interference pattern. Three coherent plane waves at a wavelength of 514 nm were symmetrically arranged in the transverse plane to produce a hexagonal interference pattern. The beams’ transverse wave vectors formed 120° angles, resulting in a lattice with periodicity of 11.2 µm. The photorefractive effect translated this optical interference pattern into a spatially modulated refractive index within the crystal, forming a stable two-dimensional lattice structure.
Figure 4.A probe beam is launched into a disordered photonic lattice, which is periodic in the transverse dimensions and invariant along the propagation direction (). The photonic lattice features a triangular arrangement with a periodicity of 11.2 µm and a refractive index contrast of approximately [217].
To enhance the photorefractive response, a spatially incoherent background beam of matching wavelength was used for uniform illumination. The background intensity was tuned to equal the maximum intensity of the lattice-forming beams, ensuring optimal refractive index modulation. A bias electric field of approximately 2000 V/cm was applied to the crystal to control the modulation depth, achieving a refractive index contrast of approximately . Random perturbations were superimposed onto the periodic lattice by adding a speckled light field. This field was generated by passing a Gaussian laser beam through an axicon (a conical lens) to create a Bessel beam. The Bessel beam was then transformed into a spatially incoherent speckle pattern by a rotating diffuser. The diffuser introduced random phase and amplitude variations, which were subsequently combined with the periodic lattice-forming beams. The speckled beam and lattice-forming beams were carefully aligned to ensure coherence, allowing their combined optical field to maintain a stable structure along the -axis. The result was a triangular lattice with controlled disorder, invariant in the propagation direction, a critical condition for observing transverse localization.
The degree of disorder was adjusted by varying the intensity of the speckled beam relative to the periodic beams. This ratio defined the balance between the random and ordered components of the refractive index modulation. Importantly, the total power of the combined beams was kept constant, ensuring consistency in the overall light intensity while precisely tuning the disorder level. To explore the statistical properties of Anderson localization, multiple realizations of the disordered lattice were generated. The rotating diffuser allowed the speckled pattern to be varied systematically, producing new configurations of disorder while preserving their statistical properties. Each realization was maintained for a duration sufficient for the photorefractive crystal to stabilize (approximately 2 min), and the crystal was reset between experiments by illuminating it with uniform white light to erase residual patterns.
A Gaussian probe beam, also at 514 nm, was injected into the disordered lattice to study its propagation dynamics. The probe beam’s initial width (full width at half-maximum) was 10.5 µm, allowing for clear observation of its diffraction, diffusion, or localization behavior. Statistical measurements were conducted by recording the probe beam’s output intensity profile for numerous disorder realizations. This systematic approach enabled the characterization of the transition from ballistic transport to diffusive spreading and, ultimately, to Anderson localization.
3.5 Implementation of Non-Hermitian Physics in Optical Microcavities
Using non-Hermitian physics to enhance optical sensing introduces a powerful method for detecting minute environmental perturbations with exceptional sensitivity. Chen et al.[223] reported an experimental realization of an EP-enhanced optical sensor, demonstrating significantly boosted sensitivity near the EP. The core of their system is a silica microtoroid resonator, which is evanescently coupled to a tapered optical fiber for both input and output of light (see Fig. 5). Owing to its circular symmetry, the microtoroid supports degenerate clockwise (CW) and counterclockwise (CCW) whispering-gallery modes (WGMs), each with identical resonance frequencies but orthogonal eigenstates—characteristics of a diabolic point (DP). To transition the system from a DP to an EP, the authors introduced two nanoscale silica tips into the resonator’s mode volume. Acting as Rayleigh scatterers, these tips mediate backscattering between CW and CCW modes, enabling controllable coupling.
Figure 5.Experimental setup for the study of exceptional-point sensors[223]. (a) A tunable laser diode (TLD) injects probe light into a whispering-gallery-mode (WGM) microtoroid resonator via a fiber-taper waveguide. The injection direction—CW or CCW—is selected using an optical switch (OS). Fiber polarization controllers (PCs) are employed to optimize the coupling efficiency. Transmitted and reflected signals are detected at the output ports by photodetectors (PDs) and recorded through an oscilloscope (OSC). The relative positions of the resonator and nearby silica nano-tips are precisely controlled using nano-positioning stages. Additional components include an optical circulator (C). (b) Optical microscope image showing the microtoroid resonator coupled to the fiber-taper and surrounded by three silica nano-tips. (c) Scanning electron microscope (SEM) image of the microtoroid, with measured major and minor diameters of approximately 80 and 100 µm, respectively.
The experiment proceeded by first identifying a resonance mode in the transmission spectrum that exhibited no detectable splitting, indicating the absence of intermodal coupling and hence operation at the DP. The appearance of a reflection signal upon the introduction of the first scatterer confirmed the onset of modal interaction and mode splitting. Subsequently, a second nano-scatterer was positioned with high precision, and its effective interaction strength was carefully tuned to bring the system to the EP condition, where both the eigenfrequencies and eigenvectors coalesce. This EP configuration forms the basis of their sensing mechanism.
This cavity-based experimental setup offers significant advantages for practical optical sensing applications[90,224–227]. The compact size of the microtoroidal cavity and its high quality factor allow for strong light-matter interactions in a confined space, enabling the detection of nanoscale particles, biomolecular interactions, or subtle environmental changes. The modularity of the system, including the precise positioning of scatterers and tunable gain-loss balance, provides flexibility for tailoring the sensing performance to specific tasks. Additionally, the platform’s ability to amplify weak perturbations makes it suitable for applications in environmental monitoring, chemical detection, and biomedical diagnostics. By leveraging the intrinsic properties of the cavity and the physics of EPs, this experimental system paves the way for the development of highly sensitive and robust optical sensors.
In addition to providing a flexible platform for simulating non-Hermitian models, photonic systems also stand to directly benefit from the unique physical features enabled by non-Hermitian engineering. One notable advantage is the ability to control gain and loss in a highly tunable manner, which opens up new avenues for photonic device functionalities beyond the reach of conventional Hermitian systems. For example, non-Hermitian physics opens new avenues for realizing mode-tunable lasers by the manipulation of coupling terms between cavities[173,174], optical switches[175], and reconfigurable topological photonics[176,177].
Different photonic platforms provide distinct mechanisms for implementing non-Hermitian effects, each with varying trade-offs in terms of scalability, control, and robustness. Table 1 summarizes and compares several representative platforms, including bulk optics, time-multiplexed systems, waveguide arrays, photonic lattices, and microcavity-based setups. Key metrics such as gain-loss mechanism, system size, operational timescale, and disorder sensitivity are highlighted to illustrate the diversity and potential of current experimental approaches.
Table 1. Comparison of Key Metrics Across Different Photonic Platforms for Non-Hermitian Implementations.
Table 1. Comparison of Key Metrics Across Different Photonic Platforms for Non-Hermitian Implementations.
Platform
Typical gain-loss mechanism
System size (sites or steps)
Timescale (per step or cavity round trip)
Disorder sensitivity
Bulk optics (Sec. 3.1)
Passive photon loss (e.g., PPBS)
steps
(per step)
Sensitive to optical alignment
Time-multiplexed systems (Sec. 3.2)
Polarization-dependent loss in loops
steps
(per loop cycle)
Robust to alignment errors
Waveguide arrays (Sec. 3.3)
Controlled propagation loss or detuning
Up to waveguides
Propagation length scale
Static disorder strongly affects localization
Photonic lattices (Sec. 3.4)
Photorefractive index modulation
2D lattices with sites
Continuous propagation; quasi-static
Adjustable disorder level
Microcavities (Sec. 3.5)
Intrinsic loss + engineered scatterers
Single or coupled resonators
Roundtrip time
Sensitive to fabrication defects
4 Introduction to Non-Bloch Band Theory via Photonic Quantum Walk
In Hermitian systems, Bloch band theory provides a foundational framework for describing the bulk properties of periodic structures. Bloch’s theorem ensures that the eigenstates of a particle in a periodic potential can be written as plane waves modulated by lattice-periodic functions. This leads to the formation of energy bands in momentum space, defined over the Brillouin zone. Importantly, topological phases and their associated edge states can be understood through the bulk-boundary correspondence, which links bulk topological invariants to the existence of boundary modes. However, in non-Hermitian systems—where energy non-conservation arises from gain, loss, or asymmetric couplings—this conventional framework breaks down. A prominent example of this breakdown is the non-Hermitian skin effect (NHSE)[66], where a macroscopic number of bulk eigenstates become localized at the boundaries of a finite system. This phenomenon defies the standard assumption that bulk states are extended, rendering Bloch’s theorem inapplicable under open boundary conditions.
Due to the NHSE, the traditional bulk-boundary correspondence fails: the energy spectra and topological edge states observed in systems with open boundaries cannot be predicted from the Bloch band structure computed under periodic boundaries. To address this issue, the concept of the generalized Brillouin zone (GBZ) was developed. Instead of defining the band structure over real wavevectors, the GBZ uses complex-valued quasi-momenta to capture the true spatial behavior of eigenstates in non-Hermitian systems. Non-Bloch band theory arises from this framework. It generalizes the conventional band theory by constructing energy spectra over the GBZ, ensuring consistency between bulk and boundary properties. This theory has successfully explained the spectral features and topological characteristics of various non-Hermitian systems, including photonic lattices, where gain and loss are naturally present.
In this section, we illustrate the non-Hermitian skin effect, the generalized Brillouin zone, and the breakdown of conventional bulk-boundary correspondence through a quantum walk model. These concepts form the foundation of non-Bloch band theory. We then discuss the applications of this theory in explaining non-Bloch EPs, measuring the non-Bloch invariant, and understanding non-Hermitian edge bursts.
4.1 Skin Effect in Non-Hermitian Quantum Walks
In this section, we discuss the skin effect, which is the most significant effect in non-Hermitian quantum walks. To do so, we derive a formal expression for the effective Hamiltonian in momentum space. Let us look at a quantum walk model as where are two shift operators, and is the gain-loss operator. The Floquet operator is non-unitary due to the gain-loss operator with non-zero . We start from the Fourier component of in momentum space as
Here, the 2-by-2 identity matrix is denoted by . The form of close to is identical to the term of in the non-Hermitian Su–Schrieffer–Heeger (SSH) model, where there is a non-Hermitian skin effect. This resemblance offers a natural explanation for the ’s skin effect[66].
Since the effective Hamiltonian is related to through , the right-eigenvectors are defined through as and the left-eigenvectors satisfy . Here , and . It is straightforward to derive
Then, the can be written as
Note that Eq. (45) gives the effective Hamiltonian in momentum space, which is already quite complicated. The transformation of to real space renders the problem even more intractable. Therefore, is usually not treated directly. Instead, we choose to concentrate on the Floquet operator , which is the standard procedure in the study of quantum-walk dynamics in discrete time. The chiral symmetric non-Hermitian Hamiltonian lacks a simple form, yet it has non-Hermitian skin effects and the conventional bulk-boundary correspondence is broken[62,228–231].
The derivation of the effective Hamiltonian is based on the PBCs. It is straightforward for us to check the eigenenergies corresponding to in Eq. (45) and to the Floquet operator in Eq. (39). As shown in Fig. 6(a), the red line represents the analysis result according to , while the blue dots represent the numerical results. We can observe a good agreement between these results. The corresponding eigen wavefunctions are shown in Fig. 6(a). It can be observed that for non-Hermitian quantum walks, all the eigenstates are extended into the whole lattice under the PBC. This is consistent with the description of eigen wavefunctions satisfying in the Bloch band theory, where , with representing the eigenstate. In fact, both Hermitian and non-Hermitian systems satisfy the Bloch band theory under the PBC.
Figure 6.(a) Quasi-energy spectrum of the effective Hamiltonian of non-Hermitian quantum walks under the PBC. The red lines represent the analytical results, and the blue dots represent the numerical calculation. (b) Distribution of eigenstates. The parameters selected are , , and .
Interestingly, it causes significant changes in the energy spectrum of non-Hermitian systems that change the boundary condition from PBC to OBC. As shown in Fig. 7(a), by keeping all the parameters unchanged and only modifying the boundary conditions to OBC, we observe that the energy spectrum of the non-Hermitian system no longer resembles Fig. 6(a) but instead undergoes substantial changes. The numerical results are depicted in Fig. 7(a), where almost all the eigenenergies are real numbers. Additionally, the eigenstates are no longer diffusely distributed into whole lattices but instead localized at the left boundary position, as shown in Fig. 7(b). This phenomenon is unique to non-Hermitian systems under the OBC and was called “skin effect” first in Ref. [66]. It was first observed in photonic quantum walks, and subsequently demonstrated in various experimental systems.
Figure 7.(a) Eigenenergy spectrum of non-Hermitian quantum walks under the OBC. (b) Distribution of eigenstates. For all eigenstates, the distributions are localized near the boundary. The parameters selected are the same as those under the PBC.
In Ref. [232], the non-Hermitian skin effect was experimentally observed in discrete-time non-unitary quantum walks of single photons, providing compelling evidence of this exotic phenomenon in photonic systems. The authors implemented the Floquet operator featuring polarization-dependent loss, thereby simulating non-Hermitian dynamics with tunable gain-loss parameter . Using a domain-wall configuration on a 1D lattice with site-dependent coin parameters, they demonstrated that for finite , all bulk eigenstates became exponentially localized near the boundaries—hallmarks of the non-Hermitian skin effect. Experimentally, the setup consisted of a single-photon interferometric network where the photon’s polarization encoded the coin state, and spatial modes encoded the walker’s position. A partial measurement operator selectively attenuated the coin state, effectively introducing controlled non-unitarity. The skin effect was observed in a seven-step quantum walk, with photon populations measured after each step. For , photon distributions spread ballistically over time [Figs. 8(a)–8(d)], while for , a pronounced localization at the boundary was observed [Figs. 8(e)–8(h)], independent of the initial coin state and in the absence of topological edge states—thereby eliminating topological contributions as the cause of localization.
Figure 8.(Upper row) Spatial probability distributions of a seven-step unitary quantum walk with and different initial states . Both the time-dependent probability distribution (a), (c) and the distribution at the last step (b), (d) are shown. (Lower row) Spatial probability distributions of a seven-step non-unitary quantum walk with and different initial states . Both the time-dependent probability distribution (e), (g) and the distribution at the last step (f), (h) are shown[232].
The skin effect was further supported by numerical simulations and comparison with analytical predictions using generalized Brillouin zones, where bulk eigenstates deviate from Bloch wave behavior. This work provides a direct experimental manifestation of the non-Hermitian skin effect in a quantum photonic system and lays the groundwork for deeper exploration of non-Hermitian topological phases in open systems.
4.2 Bulk-State Wave Functions and Generalized Brillouin Zones
The effective Hamiltonian presented above is valid under the PBC, but the results no longer hold for the OBC. Under PBC, the distribution of both eigenvalues and eigenstates can be described using the Bloch band theory. However, if the boundary conditions change, such as open boundaries or domain wall boundaries, we find that all bulk states, except for the boundary states, are localized near the boundary, resulting in the presence of the skin effect. The Bloch band theory cannot provide a reasonable explanation for the skin effect. In other words, when the skin effect is present, the Bloch band theory fails.
To explain this phenomenon, researchers have generalized the Brillouin zone to the GBZ and then proposed the non-Bloch band theory. In this section, we show the calculation of GBZ in our non-Hermitian quantum-walk system. We closely follow the non-Bloch band theory in Ref. [66], where similar quantities for a static non-Hermitian SSH model is derived. Whereas characterizing quantum-walk dynamics is more complicated due to an enlarged parameter space, the general recipe remains the same: (i) write down the ansatz wave function of the bulk states, whose variational parameters , which is the spatial-mode functions, underly non-Hermitian skin effects; (ii) derive the set of linear equations from the eigen-equation of (or the Schrödinger’s equation in the static case), both in the bulk and at boundaries; (iii) send the coefficient matrix of the linear equations to zero in the thermodynamic limit, which allows for the solution of the eigenspectrum of , as well as spatial-mode functions; (iv) find the non-Bloch topological invariants and generalized Brillouin zones. To derive the GBZ under the OBC, we consider the example with the Floquet operator as Eq. (39). It is possible to split into . Under the OBC, both and are constants in the bulk. Therefore, the Floquet operator can be rewritten as where and for . The method used to calculate the GBZ is based on Ref. [66]. Generally, the eigenstates of have the form where is a function of spatial positions and is the corresponding coin state. According to the eigen equation, we have the eigen equation of the bulk as
It is obvious that Eq. (48) has two solutions, which we denoted as and . Moreover, in the thermodynamic limit, the OBC implies a GBZ equation as
Combining this GBZ equation with the quadratic equation in Eq. (48), we have
Therefore, the GBZ is a circle in the complex plane as , with , where the radius is . As shown in Fig. 9(a), in the complex plane, the red solid line represents the GBZ, while the black dashed line represents the unit circle, which corresponds to the conventional Brillouin zone (BZ).
Figure 9.(a) BZ and GBZ in the complex plane. The red solid line represents the GBZ, while the black dashed line represents the BZ. The red circle is smaller than the black circle, indicating that all bulk states are localized at the left boundary position. (b) The eigenenergy spectrum under the GBZ. (c) The analytical results of the eigenstates, which are localized at the left boundary position. The parameters selected are , , and .
Based on the GBZ, we can obtain the effective Hamiltonian under open boundary conditions by a simple substitution relationship, i.e., replacing in the PBC with . Simultaneously, it is straightforward to derive the energy spectrum under open boundary conditions as
The analytical results are shown in Fig. 7(b), which are consistent with the numerical calculations in Fig. 7(a). Furthermore, we can also deduce the spatial distribution of the eigenstates, denoted as , as depicted in Fig. 7(c), which agrees well with the numerical calculations in Fig. 7(b). From the expression of , we can conclude that if , as increases, gradually decreases, exhibiting a skin effect at the left boundary. If , gradually increases, exhibiting a skin effect at the right boundary. The aforementioned findings completely show that the GBZ theory explains the skin effect seen in non-Hermitian quantum walks. The GBZ-based non-Bloch band theory is essential to the advancement of non-Hermitian physics.
4.3 Non-Bloch Winding Numbers in Different Time Frames
In this section, we introduce a different method for determining the non-Bloch topological invariants. The method entails calculating the non-Bloch winding numbers for the Floquet sequence at various time intervals. This method has been used in the Hermitian case to compute Bloch topological invariants. In this instance, the periodized Floquet operators in Ref. [231] yielded the same topological invariants and as those found below.
Initially, we show the computation of the winding numbers of the quantum walk in the time interval denoted by . The components of in the momentum space of the two bulks are obtained from Eq. (42) as where and . As is customary for determining the Bloch winding numbers, we first transform by unitary means, . This provides where . Following that, the generalized Zak phase defines the Bloch winding number as
Here, and are the right and left eigenstates of , respectively. More specifically, these states are solutions of eigen equations where the eigenvalue is given by . Therefore, we have the Bloch winding number as
In contrast, bulk states localize when there is a non-Hermitian skin effect. In this instance, the non-Bloch winding numbers must be determined while taking the GBZ into account. In light of this, we rewrite as , where corresponds to the modified quasi-momentum in the th GBZ of the corresponding bulk. The non-Bloch winding numbers can be obtained by replacing in Eq. (56) by . Therefore, we have
In Eq. (57), the integration is taken over the th GBZ. After numerical calculations, we found that are the same for the same bulk with integration over different GBZs of the bulk. Therefore, the index of GBZ in the left-hand side of Eq. (57) can be dropped. Using the method previously described, we may determine these topological characteristics in a different Floquet operator time frame as
There are two non-Bloch topological invariants for , which are defined as where denotes the non-Bloch winding numbers of the two bulks. Fig. 10 displays the topological invariants that were computed. The outcomes are identical to those found in Ref. [232] when using periodized Floquet operators with branch cuts. Via the non-Hermitian bulk-boundary correspondence, these topological invariants successfully predict the existence of topological edge states as well as the number of these states.
Figure 10.Non-Bloch topological invariants are calculated with Eq. (59). Non-Bloch topological invariants (a) and (b) . The red (blue) solid line represents ().
Building on the theoretical framework of non-Bloch band theory, Ref. [233] demonstrated the experimental detection of non-Bloch topological invariants in a single-photon discrete-time quantum walk. Utilizing two complementary schemes, they accessed these invariants via bulk quantum dynamics, independent of boundary effects. In the first method, the biorthogonal mean chiral displacement was extracted by reconstructing left and right quantum states from time-resolved tomography, allowing direct inference of the non-Bloch winding number when the quasienergy spectra remained entirely real. Experimental results showed excellent agreement with theoretical predictions across multiple topological phases [Fig. 11(c)], confirming phase transitions through quantized jumps in . In a second approach, the team implemented a quench protocol between different Floquet operators and , mapping the spin texture dynamics onto the generalized Brillouin zone. The resultant spin vector field exhibited skyrmion-like structures in the space when the non-Bloch winding numbers of and differed, a hallmark of topological distinction. The quantity , which quantifies the contrast in spin dynamics at and , served as a sensitive observable to detect non-Bloch phase transitions [Fig. 11(d)]. Both detection methods remained robust under static and dynamic disorders, demonstrating the intrinsic and stable nature of non-Bloch invariants. This work thus confirms the non-Bloch winding number as a measurable bulk property, firmly establishing non-Bloch topological invariants as essential tools for characterizing non-Hermitian topological systems.
Figure 11.Detecting non-Bloch winding numbers[233]. Topological phase diagrams under the (a) non-Bloch and (b) Bloch band theories. Different topological phases are characterized by the non-Bloch (Bloch) winding numbers as functions of the coin parameters . (c) The measured average chiral displacement (blue symbols) for six-step quantum walks. (d) The measured (blue symbols) along the blue path for six-step quantum walks.
4.4 Non-Bloch Bulk-Boundary Correspondence for Edge States
The bulk-boundary relationship has attracted a lot of attention in condensed matter physics. The relationship between a bulk system’s characteristics and the behavior of its edge or border states is referred to as this correspondence. Determining and forecasting the topological characteristics of materials requires an understanding of this relationship. As a result, research on non-Hermitian bulk-boundary communication has been increasingly interesting in recent years. Non-Hermitian systems have complicated energy spectra and defy the bulk-boundary connection, in contrast to ordinary Hermitian systems. Exploring the unique topological features of non-Hermitian systems requires a thorough understanding of the link between bulk properties and edge state behavior. We compute the non-Bloch topological invariants based on the GBZ. The outcomes exactly match the topological edge modes, which represent the non-Bloch bulk-boundary correspondence, as shown in Fig. 12.
Figure 12.Results of non-Bloch topological invariants. (a) Quasienergy spectrum’s absolute values for various . (b) Winding number differences between the two bulks for the - and zero-modes (blue and red, respectively). , , , and are the parameters.
The generalized bulk-boundary correspondence in non-Hermitian systems was experimentally validated via discrete-time quantum walks of single photons with polarization-dependent loss. By employing a time-integrated state reconstruction scheme, the authors distinguished topological edge states from skin-localized bulk states. The experimentally observed edge states at and matched the predictions of non-Bloch winding numbers computed over generalized Brillouin zones, while conventional Bloch invariants failed.
These results unambiguously confirm the non-Bloch formulation of bulk–boundary correspondence in photonic systems[232]. In this work, discrete-time non-unitary quantum walks of single photons are employed to experimentally verify this generalized correspondence. Both the skin effect and the emergence of topological edge states with quasi-energies and are observed. Crucially, the presence of these edge states aligns with predictions from non-Bloch topological invariants, while conventional Bloch invariants fail to capture them. This study establishes non-Hermitian bulk-boundary correspondence as a fundamental principle and highlights its importance for understanding topology in open quantum systems.
4.5 Extensions of Non-Bloch Band Theory: Non-Bloch Symmetry and Edge Burst
Non-Bloch band theory has emerged as a pivotal advancement in the study of non-Hermitian systems, offering a consistent and predictive framework where conventional Bloch theory fails due to the NHSE. By extending the BZ to complex momenta, the theory enables accurate characterization of bulk spectra, topological invariants, and boundary phenomena under open boundary conditions. This formalism has been instrumental in elucidating a broad range of non-Hermitian phenomena, including generalized topological phase transitions, non-Bloch symmetry protection, and dynamical boundary effects. Of particular interest are two recent directions: the development of non-Bloch symmetry, where NHSE modifies the spectral structure of parity-time symmetric systems; and the emergence of the non-Hermitian edge burst, which was first proposed in Ref. [234], a real-time boundary-localized dissipation process requiring both NHSE and imaginary gap closure. These phenomena have been experimentally verified in photonic quantum walks, revealing the broad applicability of non-Bloch theory to both spectral and dynamical regimes.
Xiao et al.[203] reported the first experimental observation of non-Bloch symmetry using discrete-time photonic quantum walks. Their system, consisting of a Floquet operator with tunable loss and coin operations, realized a domain-wall geometry between two topologically distinct non-Hermitian regions. By varying the coin angle , they observed a transition from exact to broken phases, determined via the corrected photon survival probability , which grows exponentially in the broken phase. The corresponding quasienergy spectra under open boundary conditions were computed using the GBZ and exhibited exceptional points not captured by Bloch theory. They show the stark contrast between the imaginary parts of the non-Bloch and Bloch spectra, while present the measured time-resolved probabilities, highlighting the transition near the predicted critical angle . The experiment features a single-photon interferometric network implementing rotations, displacements, and polarization-dependent loss via beam displacers and PPBSs.
Another extension is non-Hermitian edge burst. It refers to the dramatic accumulation of dissipation at a boundary during time evolution, driven by the NHSE and imaginary gap closure. Xiao et al.[204] demonstrated this effect in a photonic quantum walk with a domain wall separating two regions with different coin parameters. In configurations where the imaginary gap of the effective non-Hermitian Hamiltonian closes, loss initiated deep in the bulk accumulates at the boundary rather than dissipating locally. This behavior was captured by measuring the space-time-resolved loss and the total accumulated loss , with a prominent boundary peak appearing only when the imaginary gap vanishes. They confirmed the presence of a bulk-edge scaling relation , indicating increasing burst intensity with distance from the boundary.
Zhu et al.[235] further verified the universality of non-Hermitian edge burst in a photonic implementation of a non-Hermitian SSH model using time-domain quantum walks. Here, lattice sites were encoded in time bins of a recycled single-photon pulse in a fiber loop. Dissipation was applied selectively to superposition states using a customized polarizing beam splitter (CPBS), while hopping between time bins was controlled via Mach–Zehnder interferometers and electro-optic modulators. The experiment had clear control over intracellular and intercellular couplings. They observed boundary-localized dissipation only when the hopping parameters satisfied , consistent with imaginary gap closure. The dissipation profile revealed a sudden jump at the boundary in Fig. 13(c), in contrast to the smooth decay seen when the gap remained open [Fig. 13(d)]. A quantitative match with numerical simulations and a scaling relation confirmed the theoretical predictions, establishing non-Hermitian edge burst as a general consequence of non-Bloch dynamics in open systems.
Figure 13.Experimental measurements of photon dissipations[235]. (a), (b) Photon dissipation probabilities. (c), (d) Accumulated dissipations that correspond, respectively, to (a) and (b), with experimental (theoretical) results drawn in red (blue).
tNon-Hermitian photonic systems have revealed a rich landscape of emergent quantum phases that transcend traditional Hermitian classifications. Beyond the non-Hermitian skin effect and generalized band theory, recent efforts have focused on realizing and characterizing novel non-Hermitian analogs of topological insulators, gapless semimetals, and even quasicrystals. These systems exhibit unique spectral structures, non-Bloch topology, and enhanced sensitivity to boundary conditions, often governed by complex energy band topology and symmetry-protected non-Hermitian features. In particular, non-Hermitian topological insulators host boundary modes fundamentally distinct from their Hermitian counterparts, while non-Hermitian semimetals feature exceptional points and rings with nontrivial vorticity. Meanwhile, non-Hermitian quasicrystals exhibit fractal spectra and critical localization properties that challenge conventional paradigms. The following sections explore these emergent phases in detail, highlighting recent experimental and theoretical developments in photonic platforms.
5.1 Topological Insulators
Topological insulators are a class of materials distinguished by their unique electronic properties[60,61,236–238], where they act as insulators in their bulk while supporting conducting states at their surfaces or edges. These surface or edge states are protected by topological invariants, meaning they are robust against disturbances such as impurities or structural imperfections, as long as the perturbations do not close the energy gap. The concept of topological insulators derives from the quantum Hall effect, which demonstrated that electronic properties could be governed by topological rather than geometric or material-specific details[239–241]. This led to the broader realization that similar states could exist in materials without an external magnetic field, characterized instead by strong spin-orbit coupling[242–245]. These materials, termed “topological insulators”, exhibit a quantized conductance on their edges in two-dimensional (2D) systems or surface states that are immune to backscattering in three-dimensional (3D) systems. The theoretical foundation of topological insulators involves the concept of band structure topology, described by the TKNN number[246] or Chern number[247] in 2D systems. This topological invariant determines the number of edge states and their direction of propagation. In 3D systems[248–250], the topological classification is more nuanced and can involve topological invariants, which dictate the presence of an odd or even number of Dirac cones on the material’s surface.
These topological characteristics are inherently linked to the symmetries of the system, particularly time-reversal symmetry, which plays a crucial role in classifying topological insulators. The robustness of surface states is due to the fact that any local perturbation cannot couple states at the same energy with opposite momenta, thereby preventing backscattering. Non-Hermitian topological insulators extend the concept of topological phases beyond the conventional Hermitian framework, addressing systems where gain, loss, or non-reciprocal interactions play crucial roles. While traditional topological insulators are defined in terms of Hermitian Hamiltonians, ensuring that the energy spectra are real and the eigenstates are orthogonal, non-Hermitian systems introduce complex spectra and non-orthogonal eigenstates, leading to phenomena like the non-Hermitian skin effect, where eigenstates localize at the boundaries of the system.
The exploration of non-Hermitian topological insulators often focuses on how their topological invariants must be redefined to account for the complex energy values and the altered bulk-boundary correspondence. In non-Hermitian systems, the usual bulk-edge correspondence breaks down because the topological invariants calculated under periodic boundary conditions do not necessarily predict edge states under open boundary conditions, a cornerstone issue that requires the introduction of new theoretical tools like the generalized Brillouin zone. Recently, a study[251] focused on a variant of the SSH model adapted for non-Hermitian systems. The Hamiltonian for this model, as illustrated in Fig. 14, incorporating both non-Hermiticity and disorder, is expressed as where () and () are creation (annihilation) operators for the lattice sites and in the th cell, respectively. Here, denotes the disorder-dependent intercell hopping amplitude, and and are the non-Hermitian, directionally dependent intercellular hopping amplitudes.
Figure 14.Sketch of the non-Hermitian disordered SSH model[251]. The dotted box represents the unit cell, with and indicating Hermitian intercellular and nonreciprocal intercellular hoppings, respectively.
In contrast to the site-potential disorder, the pure tunneling disorder is crucial for preserving the chiral symmetry. In particular, we consider the hopping terms as where and are the intracellular and intercellular tunneling energies, and are the independent random numbers chosen uniformly in the range , and and are the disorder strengths. We hereafter assume the nonreciprocal term with the non-Hermiticity parameter .
This non-Hermitian generalization of the SSH model introduces unique phenomena such as the non-Hermitian skin effect and allows for the emergence of non-Hermitian topological Anderson insulators. These phases are characterized by the interplay of non-Hermitian parameters, such as gain and loss, with disorder-induced localization effects. In non-Hermitian systems, the conventional bulk-boundary correspondence is modified, requiring an adjustment to the definitions of topological invariants. The generalized open-boundary winding number is calculated using the relation where is the chiral operator, is the flat band Hamiltonian under open boundary conditions transformed via a similarity transformation that preserves the chiral symmetry, and is the position operator. Here, is the effective system length considered in the trace operation, and denotes the trace over a specified subsystem length. This winding number is applicable for systems characterized by non-Hermitian properties and disorder, as it effectively encapsulates the open boundary conditions typical in non-Hermitian topological phases. The disorder-averaged winding number is defined as where is a practical number of configurations sufficient for averaging. Figure 15(a) illustrates the disorder-averaged winding number as a function of and disorder strengths and . It is observed that as increases from 0 to 3.5, the topological phase with extends from a region where to a larger region where . This indicates that nonreciprocal hopping—incorporating gain and loss mechanisms—can induce a topological phase when the corresponding Hermitian limit () is a trivial or critical phase, particularly when . The outcomes for the critical scenario of are depicted in Fig. 15(b) using . Furthermore, Fig. 15(b) demonstrates that the combination of non-Hermiticity and disorder can engender a topological phase.
Figure 15.(a) Disorder-averaged winding number as a function of non-Hermiticity and disorder strength [251]. . (b) as a function of and for a nonreciprocal term , with . Other parameters are , .
The interplay of disorder, non-Hermiticity, and topology has significant implications for the localization behaviors within quantum systems. This investigation utilizes a non-Hermitian quantum walk in a fiber network to simulate a non-Hermitian topological Anderson insulator[206]. The system’s dynamics are governed by the Floquet operator defined as Eq. (39). The non-Hermitian skin effect is primarily evidenced by examining the bulk dynamics through the Lyapunov exponent
This measurement quantifies the directional propagation of eigenstates, indicating the presence of the non-Hermitian skin effect. A finite value of the Lyapunov exponent at a non-zero shift velocity highlights the impact of the non-Hermitian skin effect, contrasting with its symmetric profile in the absence of the non-Hermitian skin effect. The implemented 10-step quantum walk experiments measure polarization-averaged growth rates and photon probability distributions to confirm the non-Hermitian skin effect, as shown in Fig. 16. Upon introducing disorder, observations reveal that an asymmetric profile persists at low disorder strengths, with notable changes as disorder increases, demonstrating the competition between Anderson localization and the non-Hermitian skin effect as shown in Fig. 17.
Figure 16.Lyapunov exponent for bulk dynamics[206]. (a), (b) Polarization-averaged growth rates and (c), (d) polarization-resolved photon distribution for Hermitian quantum walks with (first column) and non-Hermitian quantum walks with (second column). The horizontal dashed lines indicate the threshold below which significant photon loss renders experimental data unreliable. The numerical findings for horizontal (vertical) polarizations are represented by the blue top (red bottom) parts for each bar in (c) and (d). The experimental probability distributions for vertical polarization (the total of the two polarizations) are represented by the white (black) dots.
Figure 17.Measured and polarization-resolved probability distributions with non-zero [206]. The other parameters of the quantum walks are , , and . It should be noted that the system is topologically trivial in (b) but topologically nontrivial in (a) and (c). The results are averaged over 20 disorder setups in the upper panel. The same symbols as in Fig. 16 are used here. (d)-(f) Experimental data (symbols) and numerical results (bars) corresponding to (a)-(c) in Ref. [206].
This presentation highlights the critical experimental validation of theoretical predictions concerning the non-Hermitian skin effect and its modulation by disorder in a non-Hermitian quantum walk scenario. The results underscore the complex interactions within the system that significantly influence its topological properties. Continuing with the analysis of the effects of disorder on the topological features of the system, the next focus is on the biorthogonal chiral displacement, which provides further insights into the interaction between disorder and the non-Hermitian skin effect. The biorthogonal chiral displacement is measured to elucidate the effect of disorder on the topological phase boundaries. It is calculated using where and , , the subscript denotes the th disorder configuration (with configurations in total), and is the position operator. Figure 18 presents the critical response of the system to varying levels of disorder and non-Hermiticity, particularly highlighting how these factors influence the topological nature of the quantum walk. Significantly, the experiment also probes the transitions at the topological phase boundary. By adjusting the coin operation parameters () differentially across a domain wall in the system, the experiments capture the emergence or suppression of topological edge states, as shown in Figs. 18(c) and 18(d). This setup effectively demonstrates how the non-Hermitian and disorder-driven dynamics within the system can lead to distinct topological behaviors, particularly near phase boundaries.
Figure 18.Results for characterizing topology with [206]. (a) Phase diagram characterized by numerically evaluated biorthogonal local marker, where and . The non-Hermitian topological Anderson insulator state corresponds to the yellow region with , which is labeled by NHTAI. (b) Measured chiral displacement for nine steps of quantum walks with averaged over 10 different configurations of randomly chosen within the range . The blue (red) symbol represent the result for (). The blue and red dashed lines represent numerical results averaged over 2000 random-disorder configurations. The blue and red solid lines represent numerical results for 400-step quantum walks averaged over 200 disorder configurations. (c), (d) Measured probability distributions near the boundary (the vertical dashed lines) after the final step.
These experimental findings corroborate the theoretical predictions about the non-Hermitian topological Anderson insulator, showcasing the intricate interplay between the non-Hermitian skin effect, disorder, and topology. The results not only confirm the theoretical framework but also enrich the understanding of non-Hermitian quantum phenomena, providing a robust platform for exploring novel quantum materials and devices that exploit these complex interactions.
5.2 Non-Hermitian Topological Semi-Metals
In exploring non-Hermitian systems, the integration of non-Hermitian properties into topological semimetals, specifically Weyl and semi-Dirac semimetals, has generated a series of novel phenomena that challenge traditional understanding in condensed matter physics. These systems are not merely theoretical curiosities but stand at the forefront of research due to their unconventional electronic properties and potential for new types of electronic devices. Non-Hermitian semi-Dirac semimetals and Weyl semimetals each exhibit unique dispersion relations and topological characteristics, yet they share foundational principles that govern their electronic behaviors and responses to non-Hermitian terms. Non-Hermitian Weyl semimetals are marked by their Weyl points where conical dispersions occur and are robust against perturbations due to their topological nature. The introduction of non-Hermitian terms such as gain and loss can lead to phenomena like the non-Hermitian skin effect, significantly altering the localization of states near boundaries[66]. On the other hand, semi-Dirac semimetals, characterized by the coexistence of linear and quadratic dispersions in orthogonal directions, demonstrate a mixed response to non-Hermitian perturbations[252].
Non-Hermitian Weyl semimetals are characterized by their unique responses to non-Hermitian perturbations, which lead to phenomena such as the non-Hermitian skin effect and alterations in the conventional bulk-boundary correspondence[59]. In these materials, the presence of non-Hermitian terms leads to a breakdown of the usual Bloch bulk-boundary correspondence, necessitating the use of non-Bloch Chern numbers to describe their topological properties accurately[66]. This modification arises due to the exceptional points in the spectrum that profoundly affect the Fermi-arc edge modes, leading to unidirectional edge motion under certain conditions[253,254]. Research into non-Hermitian semi-Dirac semimetals reveals an intriguing blend of linear and quadratic dispersions along different momentum directions, influenced by non-Hermitian terms such as gain and loss[3]. These characteristics introduce modifications that lead to new topological phases characterized by the presence of exceptional points. Unlike traditional semimetals, these materials exhibit a mixed response to non-Hermitian perturbations, where some classes exhibit the non-Hermitian skin effect while others do not, depending on the asymmetry in their hopping terms.
Both systems underline a crucial aspect of non-Hermitian physics—the profound interplay between topology and non-Hermiticity, which can alter fundamental properties like phase transitions and edge state dynamics[254]. This interaction leads to new physical insights and potential applications in devices where gain and loss can be controlled and utilized, marking a significant departure from traditional Hermitian systems in condensed matter physics. These studies pave the way for further theoretical and experimental investigations into non-Hermitian topological semimetals, promising advances in understanding topological materials under non-equilibrium conditions[254]. The insights from these investigations could lead to the development of novel quantum materials with engineered non-Hermitian properties tailored for specific technological applications.
Consider a system composed of stacked 2D layers of Chern insulators, which together form a 3D Weyl semimetal[255]. The Hamiltonian describing this two-band model is given by where and denote the intercellular and intracellular hopping amplitudes, respectively. The parameter is the onsite energy, and represents the inter-layer hopping amplitude. This Hamiltonian simplifies to that of a 2D Chern insulator when the interlayer hopping amplitude vanishes, i.e., . The introduction of intra-cell hopping amplitudes in each 2D layer imparts of non-Hermitian characteristics to the system, thus describing non-Hermitian Weyl semimetals.
In the absence of the imaginary term , the Hermitian Hamiltonian exhibits two distinct phases based on the relative strength of the parameters. When , the system features two pairs of Weyl points located at where . Alternatively, another phase manifests with a single pair of Weyl points at with , given that . The 3D Weyl semimetal in this configuration breaks time-reversal symmetry but maintains inversion symmetry; thus, the Weyl points are separated in momentum space but occur at the same energy.
The topological properties of Weyl semimetals are characterized using a topological invariant known as the Chern number. The Chern number for the th isolated band is defined as where is the Berry curvature, and is the Berry connection, with being the Bloch state eigenfunction. In non-Hermitian systems, the left and right eigenstates differ, necessitating the consideration of a combination of both eigenfunctions. The Chern number is computed within the plane, treating as a parameter. The integration covers the and coordinates. The system is identified as supporting a topologically nontrivial phase with a Chern number of one when lies between . Outside this range, it enters a topologically trivial phase with a Chern number of zero. In this analysis, parameters are set such that and , yielding .
The non-Hermitian Weyl semimetal, as represented by Eq. (68), exhibits a breaking of time-reversal symmetry while adhering to pseudo-Hermiticity. Its Hamiltonian satisfies the condition where is the product of spatial inversion and . This symmetry imposes constraints on the energy spectrum, resulting in either purely real eigenvalues or complex conjugate pairs. Consequently, the eigenvalues of this Hamiltonian are expressed as
In this framework, EPs emerge corresponding to Weyl points located along the axis at where and represent the coordinates of the inner and outer EPs, respectively. This results in two pairs of EPs appearing within the range . Moreover, for , exceptional contours (ECs) develop in the plane.
By selecting specific parameter configurations, it becomes possible to control the positions of EPs along the direction and the size of ECs by varying the non-Hermiticity parameter . Figure 19 illustrates how the size of ECs expands with an increase in the strength of the non-Hermiticity parameter. Additionally, four EPs, represented as blue dots, are observed along the line.
Figure 19.Distinctive contours (ECs) present in the non-Hermitian Weyl semimetal model (NH WSM)[255]. It displays both the real (green) and imaginary (red) parts of the exceptional surfaces for cases (a) and (b) . The intersections of these surfaces correspond to the exceptional contours. It is noteworthy that, with an increase in the strength of non-Hermiticity, the size of the ECs expands accordingly. Additionally, four EPs (indicated as blue dots) form at along the axis.
The winding number serves as a vital topological invariant in non-Hermitian systems. This invariant relates to the Berry phase acquired by the eigenstates of the system and is protected by chiral symmetry. By evaluating the winding number, a comprehensive phase diagram can be established that maps NH topological phase transitions. Given that the system exhibits chiral symmetry at and that exceptional points appear along the direction, the winding number is computed for each one-dimensional chain along the direction while treating as a parameter. The resulting expression is formulated as follows: where , and represent the coefficients of terms in the Hamiltonian, respectively.
This fractional winding number is a distinctive feature of non-Hermitian systems. It has been established that this winding number closely relates to the NH generalization of the Berry phase. The structure of the non-Hermitian winding number indicates that encircling a single exceptional point with a closed loop yields a half-quantized winding number. Importantly, when a closed contour used in winding number calculations encloses two exceptional points moving in the same winding direction, the resulting winding number is (accompanied by an non-Hermitian Berry phase value of ), reminiscent of aspects of Hermitian topology. If no such encirclement occurs, the winding number assumes a trivial value of zero. Analyzing the influence of these exceptional points on the transport properties of non-Hermitian systems is critical.
In non-Hermitian systems, it has been observed that the Hall conductance deviates from quantized levels, presenting a shoulder-like profile. This phenomenon can be interpreted as a result of the presence of pairs of exceptional points. Furthermore, the non-zero imaginary component of the energy introduces a finite lifetime for the carriers, with those possessing momenta in the direction within the range contributing most significantly to the Hall conductance. The Hall conductance exhibits an inverse relationship with the parameter due to the short lifetime of the carriers. However, when the values of the carriers lie within the bounds of , there is still a notable contribution to the Hall conductance.
5.3 Non-Hermitian Quasicrystals
Quasicrystals (QCs) are a unique class of materials characterized by their long-range order without periodicity. Unlike conventional crystals, which exhibit translational symmetry and repeat their structure periodically in space, quasicrystals exhibit a form of ordered arrangement that manifests as a quasi-periodic pattern. This distinctive arrangement is typically generated by applying non-integer tiling rules, leading to aperiodic structures on a macroscopic scale. The pioneering work[256] by Shechtman et al. in 1984 demonstrated the existence of quasicrystalline materials, initially by examining aluminum-manganese alloys, which displayed a fivefold symmetry pattern not possible in classical periodic crystals. Since then, the field of quasicrystal research has expanded significantly, revealing a variety of materials exhibiting quasicrystalline properties, including metallic, polymeric, and even photonic structures[257].
The unique symmetry properties of quasicrystals give rise to fascinating electronic and optical characteristics, making them of considerable interest in materials science and condensed matter physics. The study of quasicrystals has further led to the discovery of topological phases in these materials, often linked to their unique geometrical configurations. Recent research has increasingly focused on the interplay between non-Hermitian physics and quasicrystals, which combines the fascinating topological properties of QCs with the unusual features introduced by non-Hermiticity, such as gain and loss[258]. This burgeoning area of study highlights how non-Hermitian quasicrystals can host novel topological phases that are distinct from their Hermitian counterparts.
The introduction of non-Hermitian terms into the Hamiltonian governing quasicrystalline systems allows the emergence of unique dynamical phenomena, including the non-Hermitian skin effect and EPs[66]. These characteristics have profound implications for the electronic properties of quasicrystals, influencing localization effects and phase transitions. Recent investigations have uncovered topological phase transitions in non-Hermitian quasicrystal models, particularly through an extension of the Aubry–André–Harper (AAH) model with symmetry[258]. These studies indicate that the localization-delocalization phase transition observed at the -symmetry-breaking point is inherently topological and can be characterized by a winding number. This topological invariant effectively delineates the distinct phases of the system, highlighting how these materials can transition between metal and insulating states[259]. Moreover, experimental demonstrations of non-Hermitian quasicrystals utilizing photonic systems have shown that these structures can manifest and manipulate non-Hermitian topological features[260]. Such photonic realizations of quasicrystals promise not only to deepen our understanding of topological phases in advanced materials but also to further the development of applications in photonics and beyond.
A one-dimensional photonic quantum simulation demonstrates non-Hermitian quasicrystal dynamics[261]. The Floquet operator governing the quantum walk is formulated as where represents position-dependent phase operators, with , the inverse of the golden mean. In the absence of phase operators (), the quantum walk reduces to the traditional non-Hermitian quantum walk characterized by the non-Hermitian skin effect. The non-Hermitian skin effect arises from the interplay between effective spin-orbit coupling and polarization-dependent loss. The introduction of phase operators imposes a quasiperiodic potential on the effective Hamiltonian, leading to enhanced complexity in the system. The Floquet operator effectively simulates the dynamics of a non-Hermitian AAH model, incorporating both diagonal and off-diagonal quasiperiodic disorder. For comparison, the corresponding dynamic equation for the traditional AAH model can be derived aswhere signifies the hopping rate and represents the on-site quasiperiodic potential, with being an irrational number. The infinitesimal time evolution operator is defined as , where the time interval is infinitesimal. Utilizing the Baker–Campbell–Hausdorff formula, the expression becomes which can be expanded to yield
Consider the state , where the superscript indicates the time step. After applying the operator times, the updated state becomes
It is straightforward to observe that the quantum-walk dynamics correspond to a discrete-time evolution characterized by a lattice model with quasiperiodic potential determined by . The experimental realization of the non-Hermitian dynamics is made achievable through carefully designed fiber-optic systems, which provide a medium for the propagation of photonic states. A unique feature of quasiperiodicity is the phenomenon where the eigenstates undergo a delocalization-localization transition as the parameter increases. To characterize the localization of eigenstate , the inverse participation ratio (IPR) is utilized, defined as , where represents the amplitude of the eigenstate at position . A vanishingly small IPR indicates a delocalized state, while a finite IPR signifies a localized state.
To further investigate the delocalization-localization transition, the dynamic IPR (dIPR) is measured, defined as , with representing the normalized probability distributions of the eigenstates at time . The system is characterized as delocalized when the dIPR is vanishingly small and localized when it is finite. As illustrated in Fig. 20(a), the transition occurs near , consistent with numerical results shown in Fig. 20(b). This transition point marks the emergence of a mobility edge in the eigenspectrum, beyond which some eigenstates begin to exhibit localization. The global delocalization-localization transition, where all eigenstates become Anderson-localized, is also considered. To characterize this transition, the dynamic normalized IPR (dNIPR) is measured, defined as . In contrast to the dIPR, which provides insight into the averaged IPR, the dNIPR probes the average normalized IPR. As shown in Fig. 20(a), the global delocalization-localization transition occurs near , which aligns with numerical results displayed in Fig. 20(b). The breaking of symmetry in the system is investigated through the analysis of probability distributions. In the -symmetric phase, the quasienergy remains real and the overall probabilities are typically on the order of unity. However, when symmetry is broken, the eigenenergies acquire positive imaginary components, resulting in an increase in the overall probability over time , as illustrated in Fig. 21. In the -symmetry-broken regime, some eigenstates may still exhibit real eigenenergies, while in the -symmetry-unbroken regime, all eigenstates possess real eigenenergies. Consequently, when the system transitions from the -broken phase to the unbroken phase, a global delocalization-localization transition occurs. In this transition, all eigenstates become localized in the unbroken phase. This phenomenon elucidates the concurrence of the -symmetry-breaking transition and the delocalization-localization transition.
Figure 20.Anderson localization in the non-Hermitian quasicrystal[261]. (a) Dynamic inverse participation ratio (dIPR, blue) and dynamic normalized inverse participation ratio (dNIPR, red) are plotted with the initial state set to . Solid symbols correspond to experimental results, while hollow symbols indicate numerical simulations. (b) Numerical results for the average inverse participation ratio (, blue) and average normalized participation ratio (, red) are shown for a lattice size of under the condition of .
Figure 21.-symmetry transition in the non-Hermitian quasicrystal[261]. (a) Imaginary components of the quasienergies of the non-Hermitian quantum walk are presented for a lattice size of with . The vertical dashed line at indicates the -symmetry transition point. (b) Corrected overall probabilities are illustrated for the unbroken and broken phases of symmetry. Symbols represent experimental data, while solid lines indicate theoretical predictions.
Localization refers to the phenomenon where quantum states, which would typically spread out over time due to superposition, become confined to certain regions of space under specific conditions. In quantum mechanics, this behavior is often observed in disordered systems, such as in the case of Anderson localization. It arises when the interference of wave functions leads to the suppression of diffusion, resulting in a distinct spatial distribution of states. In non-Hermitian systems, particularly those exhibiting symmetry, localization can also be influenced by the balance between gain and loss in the system. The transition between delocalized and localized states is critical for understanding the transport properties of such systems and has profound implications for the design of quantum materials and devices.
6 Application of Non-Hermitian Physics
In the early stages of research on non-Hermitian physics, the primary focus was on the novel physical insights and phenomena that this approach, distinct from traditional Hermitian-based physics, could offer. These intriguing phenomena spurred significant growth in the field, leading to the realization of non-Hermitian systems across an expanding array of physical contexts. As the field matured, attention and interest arose in exploring the potential applications of non-Hermitian physics.
The most straightforward application of non-Hermitian physics lies in efficiently simulating and implementing open quantum systems. A quintessential example of this is the realization of quantum heat engines[92,262,263], which are open quantum systems interacting with their environment. When the environmental interactions are disregarded, the engine itself behaves as a non-Hermitian system. In this chapter, we focus on the active applications of non-Hermitian systems—specifically, how they can be used to realize applications traditionally associated with Hermitian systems. We also explore the new opportunities that non-Hermitian systems bring to these applications.
6.1 Optical Funnels
Let us begin by introducing an application of non-Hermitian systems that straightforwardly leverages the “non-Hermitian skin effect” (NHSE)[66,232,264–266]. This application, a highly efficient funnel for light as shown in Fig. 22, was demonstrated by Weidemann et al.[267] in 2020. As previously discussed, the NHSE is theoretically significant, serving as a pivotal element in the ongoing debate regarding the validity of bulk-boundary correspondence in non-Hermitian topological systems[62,66,264,268,269]. This application further illustrates that the NHSE is not just a theoretical curiosity but also a critical feature enabling practical implementations of non-Hermitian systems.
Figure 22.Conceptual illustration of light funneling[267]. The lattice design directs light from any point toward an interface, dubbed funnel opening, where it is efficiently collected.
In the experiment conducted by Weidemann et al.[267], they first realize a Hermitian SSH model[270] within a photonic lattice, where each site is coupled to its neighbors with alternating yet isotropic coupling strength. To induce the NHSE, they introduce non-Hermiticity by implementing anisotropic coupling, meaning the hopping from one site to a neighboring site differs from the hopping in the reverse direction. It is noteworthy that, in this context, non-Hermicity is introduced without the more common mechanisms of gain or loss, highlighting an alternative approach to achieving non-Hermiticity.
As discussed earlier, non-Hermiticity alone does not suffice to generate the NHSE; the presence of a boundary (or interface) is also essential. This is because, under periodic boundary conditions, the eigenmodes of both Hermitian and non-Hermitian systems remain delocalized due to translational invariance. However, when an interface is introduced, a significant difference emerges between Hermitian and non-Hermitian systems. In Hermitian systems, the interface can be created by inverting the ratio of coupling strength at a specific position, resulting in a localized mode at the interface, commonly known as a topological SSH mode. In contrast, the interface in a non-Hermitian system is introduced by reversing the direction of the anisotropy at a certain position, leading to the collapse of the entire eigenmode spectrum and the exponential localization of all eigenmodes at the interface—this is the essence of the NHSE.
The manifestation of the NHSE is that all eigenmodes localize at the interface, and excitation at any position of the lattice causes light to flow towards the interface. In the context of photonics, this means that a light signal entering the lattice is guided toward the interface and remains there, irrespective of the point of entry. Consequently, the lattice exhibits a funnel-like behavior, where light localizes exclusively along the interface.
One might intuitively assume that anisotropic coupling combined with the presence of a boundary naturally and necessarily leads to a funneling effect. However, Weidemann et al. have disproven this assumption, demonstrating that the funnel effect is strongly linked to the topological properties of the underlying lattice. The localization occurs at the interface between two regions with different topological features quantified by the winding number[254]. Here, the winding number is connected to the direction of the anisotropic coupling. To further challenge the intuitive thinking, one can investigate the robustness of NHSE against disorder[5]. Disorder can induce Anderson localization[271], a long-range interference effect. Therefore, the interplay between non-Hermiticity and disorder might reveal a suppression of Anderson localization due to non-Hermitian modulation. In Weidemann et al.’s experiment, the NHSE exhibits robustness against weak disorder but disappears under strong disorder. In other words, the Anderson localization with strong disorder behaviors is the same as the Anderson localization without anisotropic coupling[196], thereby robust to non-Hermitian modulation. Thus, non-Hermiticity cannot prevent Anderson localization as an interference effect. This behavior arises because the winding number can be defined under low disorder conditions but vanishes under strong disorder[254].
6.2 Chiral State Transfer
EPs represent a unique type of degeneracy that occurs only in non-Hermitian systems and are absent in Hermitian ones. The term “point” refers to a specific location in the system’s parameter space where this degeneracy happens. In a two-level system with two controllable parameters, varying these parameters slowly along a closed loop near the EP—although encircling the EP is preferable, it is not strictly necessary—can lead to state exchange. The direction in which the loop encircles the EP determines the direction of the state transition, with the state switching in only one direction. The chirality is signaled by the different final states for different encircling directions. While this might seem to contradict the adiabatic theorem, the chiral state transfer is a result of the complex eigenspectrum in the parameter space and the path-dependent amplification of nonadiabatic couplings[272].
Experimentally, the chiral state transfer has been observed in classical systems with gain and loss[46,273], as well as in dissipative quantum systems including photons[274], superconducting qubits[85,275], trapped ions[276], solid-state spins[277], and cold atoms[278].
Here, we will provide an example to show the phenomenon of the chiral state transfer. We consider the dynamical process driven by the non-Hermitian Hamiltonian[279]where , and . The Hermitian part of the Hamiltonian is , and an EP exists at , under .
We consider in Fig. 23 some typical solutions for encircling an EP, the path in parameter space as defined in Fig. 23(a). In Figs. 23(b) and 23(c), we have chosen a CW and a CCW encircling, respectively, , . The trajectories are calculated according to where are the left eigenstates of , with , and are the eigenvalues of . The time-evolved state is given by
Figure 23.(a) Parametric path, , . Trajectories of the CW rotation (b) and the CCW rotation (c) against the eigenspectra of the non-Hermitian Hamiltonian . The red (blue) color code indicates the eigenstate with a larger (smaller) imaginary component, and hence smaller (larger) loss. The black triangles represent the initial states, the black squares denote the final states, and the black lines illustrate the evolution trajectories.
In the counterclockwise example, the solution aligns with the adiabatic prediction, and the state transitions accordingly. Conversely, in the clockwise example, a nonadiabatic transition occurs, causing the system to return to its original state. This chiral behavior, initially discussed in Ref. [280], highlights a fundamental distinction between Hermitian and non-Hermitian dynamics. In non-Hermitian systems, complex eigenvalues lead to gain or loss effects. Consequently, even a very small nonadiabatic coupling can be exponentially amplified, resulting in the dominance of the gain eigenvector. This mechanism provides an intuitive explanation for why the adiabatic theorem generally does not apply to non-Hermitian systems.
6.3 Non-Hermitian Sensing
In contemporary quantum technology, quantum metrology and sensing[281–286] stand as one of the central applications. Conventional Hermitian quantum sensors offer substantial advantages in terms of accuracy, repeatability, and precision over classical sensors for detecting weak signals. Alongside the rapid advancements of non-Hermitian physics, numerous novel sensing techniques have been proposed and experimentally demonstrated within non-Hermitian systems[3,287–291]. These non-Hermitian sensing protocols exhibit significant improvements over traditional Hermitian quantum sensing methods.
A major category of non-Hermitian sensing is based on the EP[24,37,39,223,292–298], which is unique to non-Hermitian systems[149,299,300]. This is because the eigenspectra near EP become highly susceptible to weak perturbations. It is important to note, however, that not all non-Hermitian systems possess EPs. There are also non-Hermitian sensing protocols that operate without relying on EPs[301]. Here, we introduce two types of non-Hermitian sensing protocols: those with and without EPs. For detailed discussions (or debate) on the enhancement of non-Hermitian sensing, we refer readers to Refs. [302–308].
6.3.1 EP-based non-Hermitian quantum sensing
As an example, we introduce the non-Hermitian sensor based on EPs implemented in a dissipative single-qubit open system reported in Ref. [309]. This sensor employs an effective periodically driven[310]-symmetric non-Hermitian Hamiltonian of the form where are the Pauli operators, is the coupling strength, represents the modulation frequency, and denotes the dissipation rate. In practice, the Hamiltonian used is , a passive -symmetric system with being the identity operator. The system experiences a perturbation , where and are the perturbation’s amplitude and frequency, respectively, with being the parameter of interest.
The sensor is initialized in a specific state. After evolving the system for a duration of , the response energy can be extracted from the difference . Here, and . It is straightforward to check how varies with , which reveals sharp dips near the EPs. This feature suggests high sensitivity to changes in , which is the fundamental of EP-based non-Hermitian quantum sensing.
6.3.2 Non-Hermitian quantum sensing in the absence of EP
In 2024, Xiao et al. introduced a non-Hermitian sensing technique that operates effectively even in the absence of EPs[301]. It is worth noting that the enhanced sensing is due to non-Hermiticity, a more universal trait of non-Hermitian systems that contrasts with EP[311].
The sensor in Ref. [301] is a generic non-Hermitian qubit subject to a weak external field, represented by the Hamiltonian , where is the small parameter of the external field to be estimated. In an orthonormal basis , the bare non-Hermitian Hamiltonian is with . Its eigenstates are approximately and , with an energy splitting of . Notably, this protocol of sensing does not require ; thus, it remains unaffected by the properties of EPs.
For the unperturbed case, the dynamic of the sensor system is governed by bare non-Hermitian Hamiltonian . Starting from the initial state , the time-evolved state of the system is where . Therefore, its population in the state can be expressed as
When introducing the weak external field, the parameters become dependent on . The sensitivity of the normalized population in determines the susceptibility as where and denote the responses of to changes in and , respectively. When , it is easy to see that the population exhibits rapid variations with respect to and , leading to high sensitivity to .
The sensitivity of this sensor arises from the significant impact of weak perturbations on the population ratio when is comparable to , which is termed as the population-matching condition. This population-matching condition is achieved by leveraging the inherent non-Hermiticity of the sensor and facilitates high sensitivity. A simpler instance is a pseudo-Hermitian system with real , where can be close to unity (), allowing [311]. When approaches , can be on the order of , satisfying the population-matching condition. Xiao et al. have further shown that non-Hermitian sensing configurations beyond the pseudo-Hermitian case can also be developed based on this condition[301]. The significant difference between this sensor and those with EPs[292,293] is that this sensor exploits a non-Hermitian sensing region emerging from the dynamics starting close to an eigenstate of the unperturbed system.
In contrast to a Hermitian system, of which the evolution is unitary, becomes , with . That is, . For example, in a Hermitian sensor with , the time-evolved state is given by Eq. (82), with and , resulting in a smooth sinusoidal population as This comparison clearly demonstrates the advantage of the non-Hermitian sensor, which offers enhanced sensitivity over its Hermitian counterpart.
This protocol is experimentally demonstrated using a cyclic photonic interferometer that simulates discrete-time dynamics with adjustable non-Hermiticity. The experiment shows measured susceptibilities of for a non-Hermitian case and for the corresponding Hermitian case, which demonstrates a significant advantage of non-Hermitian sensing over the Hermitian method. Moreover, the authors have encoded the signal in the setting angle of a single wave plate, which is a real physical parameter to be estimated. In this primordial practical application, the measured susceptibility is , which still shows an enhancement of susceptibility with a ratio compared to the Hermitian protocol with measured susceptibility no greater than .
6.4 Quantum Algorithms in Non-Hermitian Systems
In the studies of non-Hermitian physics and quantum computation, quantum walks play a central role. In the latter, quantum walks provide a platform to design and realize quantum algorithms, e.g., search algorithms[312,313] and universal quantum computing[314,315]. The traditional quantum walk is considered in undirected graphs[316,317], of which the Hamiltonian is Hermitian and evolution is unitary. In 2020, Wang et al. considered the quantum walk corresponding to the problem of PageRank, which is a directed graph[318]. By realizing the quantum walk on the directed graph, of which the Hamiltonian is non-Hermitian, they have demonstrated that non-Hermitian quantum walk can efficiently solve the RageRank.
The PageRank algorithm was initially introduced by Page and Brin[319] and now serves as the dominant algorithm in search engines such as Google. The problem of PageRank can be modeled as a directed graph with nodes representing websites. Considering a directed graph with nodes, it can be characterized by an non-Hermitian adjacency matrix . In this case, the evolution of the quantum walk governed by the Hamiltonian is a non-unitary evolution . In Wang et al.’s work, this non-Hermitian quantum walk is realized by the unitary-dilation approach[320,321].
To realize PageRank via quantum walk, the walker is initialized as identically distributed on all nodes to ensure no bias toward any particular node. After that, the intrinsic relative importance of each node is quantified by the maximum probabilities , where is the appropriate evolution time. The validity of these probabilities in PageRank is certified by comparing them with the classical centrality measure. In Wang et al.’s work, they have demonstrated such a quantum PageRank with two distinct directed graphs with three nodes.
7 Summary and Outlook
Originating from a mathematical concept of non-Hermiticity, non-Hermitian systems have sparked enormous attention over the past decade. In photonic systems, the controllability and extendability have accelerated the development of non-Hermitian physics. Additionally, non-Hermitian physics also leads to new technological innovations in photonic systems. Drawing insights from recent progress in this field, we have outlined the methods for implementing non-Hermiticity across various photonic platforms and explored key concepts in non-Hermitian systems.
In Sec. 2, we have outlined the fundamental concepts of non-Hermitian Hamiltonians. We begin by discussing the construction of effective non-Hermitian systems derived from the Lindblad master equation. Subsequently, we introduce the properties of non-Hermitian Hamiltonians, such as their complex spectra and biorthogonal eigenstates. The conserved quantities and their connections to symmetry or pseudo-Hermiticity are analyzed and illustrated with concrete examples. Additionally, we review the creation and applications of EPs. In Sec. 3, we present methods for simulating non-Hermitian systems across various photonic platforms, including bulk optics, optical waveguides, optical cavities, fiber optics, synthetic dimensions in photonics, and optical metamaterials. Numerous examples are provided to demonstrate the ease of controlling non-Hermiticity in photonic systems. In Sec. 4, we delve into the introduction of non-Bloch band theory via photonic quantum walk, detailing the origins of the non-Hermitian skin effect and bulk boundary correspondence. Key concepts such as the generalized Brillouin zone, generalized BBC, and non-Bloch EPs are also introduced. In Sec. 5, we explore the unique topological phases induced by complex spectra in non-Hermitian systems under the influence of disorder and external fields. Finally, in Sec. 6, we highlight several examples that bridge intriguing phenomena and key concepts in non-Hermitian physics to practical applications in photonics, including optical funnels, chiral state transfer, non-Hermitian sensing, and quantum algorithms in non-Hermitian systems.
Novel theoretical ideas combined with new experimental approaches are expected to yield surprising results in non-Hermitian photonic systems. As theoretical models continue to evolve, they are likely to uncover novel phenomena and insights that challenge traditional paradigms, deepening our understanding of non-Hermitian physics and its practical applications. On the experimental front, a central challenge in advancing non-Hermitian photonic systems lies in the scalable simulation of large-scale non-Hermitian dynamics using smaller-scale setups. Current implementations typically rely on engineering photon loss to emulate non-Hermitian effects. However, as the simulated system size increases and the physical size of the setup decreases, photon loss becomes increasingly severe, and the cumulative influence of environmental disturbances grows accordingly. These effects can rapidly destroy quantum coherence, ultimately undermining the quantum nature of the system. Achieving scalability in non-Hermitian systems while maintaining precise control over system parameters remains a major challenge and will require innovative approaches to ensure that the relevant quantum phenomena remain both accessible and tunable. In this direction, optical waveguides, optical cavities, fiber optics, and synthetic dimensions in photonics are among the promising platforms that have already been successfully applied in exploring non-Hermitian physics, providing controlled environments where the intricacies of non-Hermitian dynamics can be investigated. These platforms allow for the creation of artificial structures that mimic complex behaviors.
Another critical issue to be particularly concerned with is that the simulation of non-Hermitian systems on optical platforms tends to rely on single-particle treatments, which are inherently limited in capturing many-body effects that could emerge in more complex, interacting systems. Moreover, the coupling mechanisms in these systems are often restricted to nearest neighbors, which imposes constraints on the kinds of non-Hermitian phenomena that can be studied, particularly those involving long-range interactions or more intricate topological effects. Therefore, new technologies aimed at constructing non-Hermitian systems with multiple photons and long-range coupling are expected to significantly accelerate research in non-Hermitian physics. These technologies could lead to breakthroughs in quantum simulations, quantum computing, and even the design of advanced photonic devices. For example, implementing long-range couplings could enable the simulation of non-Hermitian systems with more complex interactions, potentially leading to the discovery of new topological phases or quantum-critical behaviors that are not observable in systems with nearest-neighbor coupling alone.
Furthermore, advancements in technologies for the detection and precise control of various forms of non-Hermiticity are crucial. Current techniques often struggle with the required level of precision needed to manipulate and measure non-Hermitian systems with high fidelity. These technologies need to evolve to allow for the accurate measurement of the spectral properties of the systems, time evolution, and response to perturbations at the level of individual particles or even many-body systems. Additionally, the ability to control and measure multi-photon systems could allow for the exploration of quantum coherence, entanglement, and other quantum phenomena within the framework of non-Hermitian physics, providing a rich avenue for future research and technological advancement. As these new technologies emerge, they will likely open up a host of exciting possibilities for both fundamental science and applied technologies in areas such as quantum sensing, communication, and metrology.
Based on both the theoretical and experimental developments on non-Hermitian physics, a natural quest is to harness non-Hermitian systems for enhanced applications, exploiting the unique properties that arise from their non-Hermitian nature. As a promising step, the spectral sensitivity of the systems near EPs to perturbations has been demonstrated in various studies[37–39], revealing their potential for ultra-sensitive measurements. This sensitivity has opened new avenues for precision sensing, as slight perturbations to the system can induce large shifts in its response, making EPs particularly attractive for sensing applications.
However, there remains considerable debate on whether non-Hermitian systems truly offer distinct advantages over their Hermitian counterparts. Some studies suggest that by incorporating non-Hermitian systems into the framework of enlarged Hermitian systems, these systems do not provide a substantial improvement in sensing capabilities[308]. Moreover, the quantum-limited SNR at EPs is found to be proportional to the perturbation itself, implying that there may be no inherent advantage over traditional Hermitian systems in terms of improving the SNR. This challenges the initial optimism surrounding non-Hermitian sensing and underscores the need for further investigation to uncover novel methods for realizing non-Hermitian-enhanced sensing with clear benefits. A particularly promising direction is to focus on how non-Hermitian physics can be applied to reduce noise while simultaneously improving the SNR, particularly by exploiting EPs to filter out unwanted noise and enhance the desired signal. Critical phenomena such as phase transitions[322], anomalous topological features[289], and the non-Hermitian skin effect[291] have also been found to enhance sensing capabilities, enabling detection of subtle changes in the environment or system parameters. These effects have shown promise in diverse areas, ranging from the detection of weak forces to the monitoring of complex systems.
Additionally, non-Hermitian systems are emerging as a powerful tool in quantum thermodynamics, where they offer the potential to improve the efficiency of heat extraction and work generation in quantum engines. Recent discoveries, such as enhanced cooling via the non-Hermitian skin effect[323], and advancements in quantum thermal engines that operate near EPs[92,93], show that non-Hermitian systems can enable more efficient energy management, which is essential for practical quantum technologies. Furthermore, novel approaches for generating entanglement based on EPs have been proposed[324], revealing that non-Hermitian dynamics may provide new tools for entanglement generation and quantum information processing. These methods hold significant potential for improving quantum communication, cryptography, and computing by leveraging the robust, high-dimensional entanglement properties that EPs enable. In addition to their applications in quantum information, the rich topological and dynamical features of non-Hermitian systems have prompted new perspectives in quantum simulation and the study of non-equilibrium systems. These advancements collectively suggest that the field of non-Hermitian physics is poised to drive transformative innovations across a range of quantum technologies, with substantial implications for both fundamental research and practical applications.
Acknowledgments
Acknowledgment. This work was supported by the National Key R&D Program of China (No. 2023YFA1406701), the National Natural Science Foundation of China (Nos. 12025401, 92265209, 92476106, 12474352, 12305008, and 12374479), the Fellowship of China Postdoctoral Science Foundation (Nos. BX20230036, BX20240065, 2023M730198, 2024M750405, and 2024M760425), and the Beijing National Laboratory for Condensed Matter Physics (No. 2024BNLCMPKF010).