Singular-Value Statistics of
Non-Hermitian Random Matrices and Open Quantum Systems
Abstract
The spectral statistics of non-Hermitian random matrices are of importance as a diagnostic tool for chaotic behavior in open quantum systems. Here, we investigate the statistical properties of singular values in non-Hermitian random matrices as an effective measure of quantifying dissipative quantum chaos. By means of Hermitization, we reveal the unique characteristics of the singular-value statistics that distinguish them from the complex-eigenvalue statistics, and establish the comprehensive classification of the singular-value statistics for all the 38-fold symmetry classes of non-Hermitian random matrices. We also analytically derive the singular-value statistics of small random matrices, which well describe those of large random matrices in the similar spirit to the Wigner surmise. Furthermore, we demonstrate that singular values of open quantum many-body systems follow the random-matrix statistics, thereby identifying chaos and nonintegrability in open quantum systems. Our work elucidates that the singular-value statistics serve as a clear indicator of symmetry and lay a foundation for statistical physics of open quantum systems.
I Introduction
The spectral statistics serve as a defining feature of chaotic behavior in quantum systems and lack thereof, and play a fundamental role in the principles of statistical mechanics [1, 2, 3]. According to a number of numerical and analytical investigations, it is widely accepted that the spectrum of a nonintegrable quantum system follows the random-matrix statistics [4], while the spectrum of an integrable quantum system follows the Poisson statistics [5]. Here, the spectral statistics of Hermitian random matrices are universally determined by the 10-fold symmetry classification [6]. These 10 symmetry classes encompass the 3 standard (Wigner-Dyson) classes by time-reversal symmetry [7, *Wigner-58, 9], 3 chiral classes by chiral symmetry [10, *Gade-93, 12, *Verbaarschot-00-review], and 4 Bogoliubov-de Gennes classes by particle-hole symmetry [6]. Time-reversal symmetry influences the correlations in the bulk of the spectra, quantified by the Dyson index . On the other hand, the impact of chiral or particle-hole symmetry arises only around the symmetric point (i.e., zero eigenvalue) and is classified by the other random-matrix index . This 10-fold symmetry classification is also fundamental in understanding mesoscopic transport phenomena including the Anderson transitions [14, *Beenakker-review-15, 16], as well as topological insulators and superconductors [17, *Ryu-10, 19, 20, 21, 22].
Recently, there has been a remarkable surge of interest in the physics of open quantum systems. As coupling with the surrounding environment is an inevitable aspect of realistic physical systems, an understanding of open quantum systems is important for quantum simulation and technology. In contrast to closed quantum systems, open quantum systems are no longer described by Hermitian Hamiltonians. In general, the dynamics of open quantum systems is described by Liouvillian superoperators that map a density operator to another density operator [23, 24, 25]. In addition, non-Hermitian Hamiltonians effectively capture the open quantum dynamics of individual quantum trajectories subject to measurements or dissipative processes [26, 27, 28], as well as the open classical dynamics [29, 30, 31]. Researchers have discovered unique phenomena that lack analogs in closed quantum systems and hence are intrinsic to open quantum systems. Prime recent examples include the entanglement phase transitions induced by the interplay of the coherent dynamics and measurement [32, 33], as well as non-Hermitian topological phases [34, 35].
Moreover, characterization of chaos in open quantum systems has also attracted growing interest [36, *Grobe-89, 38, 39, 40, 41, *Can-19JPhysA, 43, 44, 45, 46, 47, 48, *GarciaGarcia-22PRD, 50, 51, 52, 53, 54, 55, 56, *GJ-23-ETH, 58, 59, 60, 61, 62, 63]. Such dissipative quantum chaos is relevant to the development of thermodynamics and statistical mechanics in open quantum systems. As prototypes of open quantum chaotic systems, researchers have proposed random Lindbladians [38, 40, 41, *Can-19JPhysA, 46], as well as dissipative extensions of the Sachdev-Ye-Kitaev model [47, 48, *GarciaGarcia-22PRD, 53, 54, 55, 63]. Analogous to closed quantum systems, several numerical calculations have shown that the statistics of complex-valued eigenvalues in nonintegrable open quantum systems follow the statistics of non-Hermitian random matrices, while those in integrable counterparts follow the Poisson statistics for complex numbers. Notably, non-Hermiticity changes the nature of symmetry, and the spectral statistics of non-Hermitian random matrices are no longer described by the 10-fold symmetry classification, instead universally conforming to the 38-fold symmetry classification [64, 65]. These 38-fold symmetry classes are also relevant to the classification of the Anderson transitions [66, *Hatano-Nelson-97, 68, 69, 70, 71, 72, 73, 74, 75, *Luo-21B, 77, 78, 79] and topological phases [80, 65, 81, 34, 35] in non-Hermitian systems.
In this work, we investigate the spectral statistics of singular values, instead of complex eigenvalues, in non-Hermitian random matrices. For a given matrix , its singular values are defined as the square roots of eigenvalues of or . When respects Hermiticity (i.e., ), singular values reduce to the absolute values of eigenvalues and hence contain essentially the same information as eigenvalues. However, in the case of non-Hermitian matrices (i.e., ), singular values play a distinct role from eigenvalues and find broad applications in diverse areas of science. For example, singular values represent the degree of information associated with each mode and thus form the basis for principal component analysis and various tensor-network techniques. This also implies that if is interpreted as a generator of the open quantum dynamics, its singular values quantify the amount of information in the quantum operation. The singular-value decomposition is also utilized to capture the topological phenomena and bulk-boundary correspondence in non-Hermitian systems [82, *Ramos-22, 84, 85, *Brunelli-22]. Additionally, if is interpreted as a reflection or transmission matrix, the squared singular values give the reflection or transmission probability, respectively [87, 88]. The fluctuations of singular values from sample to sample, or the suppression thereof, physically manifest themselves in the universal conductance fluctuations [89, *Lee-87, 91, 92, 93].
Here, we demonstrate that the singular-value statistics provide an effective diagnostic tool for understanding chaotic behavior in open quantum systems. We establish that singular values of open quantum many-body systems, including Lindbladians and non-Hermitian Hamiltonians, follow the random-matrix statistics in the corresponding symmetry classes. Employing the Hermitization technique [69, 80, 65, 77], we completely classify the singular-value statistics of non-Hermitian random matrices across all the 38 symmetry classes, as summarized in Tables 2-5. Importantly, we reveal that the singular-value statistics exhibit unique characteristics and serve as a clearer indicator of symmetry compared to the complex-eigenvalue statistics. In contrast to Hermitian random matrices, the level repulsion of complex eigenvalues in non-Hermitian random matrices is universally cubic, regardless of symmetry [37]. Additionally, the correlations of complex eigenvalues in the bulk of the spectra depend solely on time-reversal symmetry†, and the other symmetries are irrelevant [43, 58]. By contrast, we elucidate that all the symmetries influence the bulk spectral correlations of singular values even in non-Hermitian random matrices. Furthermore, we illustrate that these distinctions can be clearly discerned through the random-matrix indices , akin to Hermitian random matrices. Consequently, these results show the significance of singular-value statistics as a powerful practical measure for characterizing dissipative quantum chaos and its absence, thereby facilitating a deeper understanding of statistical mechanics in open quantum systems.
The rest of this work is organized as follows. In Sec. II, we describe symmetry of non-Hermitian matrices. In Sec. III, we study the singular-value statistics of non-Hermitian random matrices (Table 1 and Fig. 1). Through Hermitization, we completely classify the singular-value statistics of non-Hermitian random matrices for all the 38 symmetry classes (Tables 2-5). In Sec. IV, we demonstrate that the singular-value statistics of physical Lindbladians and non-Hermitian Hamiltonians coincide with those of non-Hermitian random matrices, characterizing the dissipative quantum chaos. In Sec. V, we conclude this work. In Appendix A, we discuss the singular-value statistics of Hermitian random matrices (Table 6). In Appendix B, we analytically derive the singular-value statistics of small random matrices for all the symmetry classes. In Appendix C, we discuss the singular-value statistics of normal random matrices. In Appendix D, we develop the symmetry classification of non-Hermitian reflection matrices.
II Symmetry of non-Hermitian random matrices
We begin with the symmetry classification of Hermitian random matrices. In this work, we focus on square matrices to consider physical applications to quantum chaos. In general, Hermitian matrices are classified according to the two types of antiunitary symmetry, time-reversal symmetry
(1) |
and particle-hole symmetry
(2) |
with unitary matrices and . As a combination of time-reversal symmetry and particle-hole symmetry, we can also introduce chiral symmetry, or equivalently sublattice symmetry, by
(3) |
with a unitary matrix . These two antiunitary symmetries and one unitary symmetry form the 10-fold Altland-Zirnbauer (AZ) symmetry classification (see also Table 6 in Appendix A) [6]. Classes A, AI, and AII are the 3-fold standard (Wigner-Dyson) classes [7, *Wigner-58, 9], classes AIII, BDI, and CII are the 3-fold chiral classes [10, *Gade-93, 12, *Verbaarschot-00-review], and classes D, DIII, C, and CI are the 4-fold Bogoliubov-de Gennes (BdG) classes [6]. These AZ symmetry classes yield the 10-fold classification of Hermitian random matrices [6, 1], as well as the Anderson transitions [14, *Beenakker-review-15, 16] and topological insulators and superconductors [17, *Ryu-10, 19, 20, 21, 22].
The nature of symmetry changes in non-Hermitian matrices, leading to the 38-fold classification [64, 65]. First, the two types of antiunitary symmetry in Eqs. (1) and (2) remain to be time-reversal symmetry and particle-hole symmetry even for non-Hermitian matrices , respectively. In addition to these symmetries, we can also introduce two additional antiunitary symmetries by
(4) |
and
(5) |
with unitary matrices and . These symmetries are respectively denoted by time-reversal symmetry† and particle-hole symmetry† since they are obtained by additional Hermitian conjugation to time-reversal symmetry in Eq. (1) and particle-hole symmetry in Eq. (2), respectively [65]. While time-reversal symmetry in Eq. (1) and time-reversal symmetry† in Eq. (4) are equivalent for Hermitian matrices, this is not the case for non-Hermitian matrices.
The unitary symmetry in Eq. (3) is still symmetry and is denoted by sublattice symmetry for non-Hermitian matrices. Similarly to the antiunitary symmetries, we can also consider the Hermitian conjugate of Eq. (3) by
(6) |
with a unitary matrix . In contrast to Eq. (3), unitary symmetry in Eq. (6) is called chiral symmetry for non-Hermitian matrices. Here, while chiral symmetry and sublattice symmetry are equivalent to each other for Hermitian matrices, they are different for non-Hermitian matrices. The combination of time-reversal symmetry in Eq. (1) and particle-hole symmetry in Eq. (2), as well as the combination of time-reversal symmetry† in Eq. (4) and particle-hole symmetry† in Eq. (5), yields chiral symmetry in Eq. (6); on the other hand, the combination of time-reversal symmetry and particle-hole symmetry†, as well as the combination of time-reversal symmetry† and particle-hole symmetry, yields sublattice symmetry in Eq. (3). Chiral symmetry in Eq. (6) is also equivalent to pseudo-(anti-)Hermiticity [94, *Mostafazadeh-02-2].
Different symmetries lead to different universality classes, resulting in the 38-fold classification of non-Hermitian random matrices. Specifically, time-reversal symmetry† in Eq. (4) changes the complex-eigenvalue statistics of non-Hermitian random matrices in the bulk of the spectra [43] while the other symmetries change the complex-eigenvalue statistics at the symmetric lines or point [58]. Moreover, the 38-fold classification characterizes the universality classes of the Anderson transitions [74, 75, *Luo-21B, 77] and topological phases [65, 81] of non-Hermitian systems. In this work, we below show that symmetry determines the universal statistics of singular values in non-Hermitian random matrices and comprehensively classify them for all the 38 symmetry classes.
III Singular-value statistics of non-Hermitian random matrices
The complex-eigenvalue statistics of non-Hermitian random matrices have recently been investigated as a diagnostic of dissipative quantum chaos. Here, we study the spectral statistics of singular values in non-Hermitian random matrices. While eigenvalues and singular values contain essentially the same information in Hermitian matrices, they exhibit distinct behavior in non-Hermitian matrices. We demonstrate that the singular-value statistics exhibit the behavior that distinguishes them from the complex-eigenvalue statistics and serve as an effective measure of dissipative quantum chaos. While we focus on non-Hermitian random matrices from the Gaussian ensemble for clarity, our results should be independent of the specific details of the ensemble and universally determined by symmetry alone. In fact, we below show that the random-matrix statistics appear even in open quantum chaotic systems, suggesting the universality of the singular-value statistics (see Sec. IV for details).
In Sec. III.1, we first provide the singular-value statistics for the 9 symmetry classes where no symmetry or only one symmetry is present (Table 1 and Fig. 1). In Sec. III.2, we reduce the singular-value statistics of non-Hermitian random matrices to those of Hermitian random matrices in the corresponding symmetry classes, using the Hermitization technique. In Sec. III.3, we provide the complete classification of the singular-value statistics of non-Hermitian random matrices for all the 38 symmetry classes (Tables 2-5).
Class | Symmetry | Hermitization | ||
A | No | AIII | ||
AI | Time-reversal symmetry () | BDI | ||
AI† | Time-reversal symmetry† () | CI | ||
AII | Time-reversal symmetry () | CII | ||
AII† | Time-reversal symmetry† () | DIII | ||
AIII | Chiral symmetry | A | ||
AIII† | Sublattice symmetry | AIII AIII | ||
D | Particle-hole symmetry () | DIII | ||
C | Particle-hole symmetry () | CI |

III.1 Singular-value statistics
We first focus on the 9 symmetry classes where no symmetry or only one symmetry is present (Table 1) and provide the complete classification for all the 38 symmetry classes shortly in Sec. III.3 (Tables 2-5). Singular values of a matrix are the square roots of eigenvalues of or . As a clear difference from eigenvalues, singular values are always real and nonnegative even in non-Hermitian matrices. Then, we calculate the distributions of singular-value-spacing ratios [Fig. 1 (a, b)], in a similar manner to the spacing ratios of real eigenvalues [96, 97]. Specifically, is defined as
(7) |
for an ordered set of singular values ’s. In classes AII, AII†, and D, all the singular values are two-fold degenerate, and we calculate by identifying the two degenerate singular values with the one singular value. While we can also study the statistical distributions of singular-value spacing (see also Appendix B), the statistical distributions of are independent of the local density of singular values and hence do not require the unfolding of singular values, which facilitates the calculations for physical open quantum systems.
The singular-value-spacing ratios exhibit the three different distributions in the presence of time-reversal symmetry in Eq. (1) or time-reversal symmetry† in Eq. (4). Here, non-Hermitian matrices without symmetry belong to class A, and those with Eq. (1) [Eq. (4)] belong to classes AI and AII (AI† and AII†) [65]. Notably, time-reversal symmetry changes the level-spacing distributions even for arbitrary singular values. By contrast, while time-reversal symmetry† changes the bulk statistics of complex eigenvalues [39], time-reversal symmetry changes the eigenvalue statistics only around the real axis [98, 99, 100, 58, 101, *2023arXiv230105022B]. Thus, the singular-value statistics give a clearer measure of time-reversal symmetry in non-Hermitian random matrices. It is also notable that all the level-spacing distributions of complex eigenvalues exhibit the cubic level repulsion [37, 43, 45] in contrast to Hermitian random matrices. Conversely, the singular-value-spacing distributions exhibit the 3-fold level repulsion even in non-Hermitian random matrices,
(8) |
for
(9) |
All the other symmetries also change the level-spacing-ratio distributions of singular values [Fig. 1 (b)]. Remarkably, in classes AIII and AIII†, where chiral and sublattice symmetries in Eqs. (6) and (3) are respected by definition, the probability density is nonvanishing even for small level-spacing ratio , which implies the weaker level repulsion between singular values. Still, the level-spacing-ratio distributions deviate from the Poisson statistics
(10) |
for uncorrelated singular values. We elucidate these unconventional level-spacing-ratio distributions through Hermitization in Sec. III.2.
While some symmetry classes (e.g., classes AI and AI†) exhibit the same level-spacing distributions, they can be distinguished by the distributions of the minimum singular value [Fig. 1 (c, d)]. Here, in contrast to the spacing ratio , the minimum singular value depends on the normalization of random matrices; in our numerical calculations in Fig. 1 (c, d), we normalize such that its average is unity (i.e., ). For example, non-Hermitian random matrices in classes AI and AI† follow the same level-spacing distributions but exhibit clearly different distributions of the minimum singular value. The former is quantified by the average singular-value-spacing ratio while the latter by the average normalized variances of the minimum singular value, as summarized in Table 1. Additionally, each symmetry class can be distinguished between by the power-law decay of the probability distribution for small ,
(11) |
with
(12) |
We also note that class AIII is characterized by unique singular-value statistics, as discussed shortly. If matrices additionally respect Hermiticity or normality, they exhibit different singular-value statistics (see Appendices A and C for details).
III.2 Hermitization
We demonstrate that the singular-value statistics are captured by Hermitization. Previously, the Hermitization technique was applied to the complex-eigenvalue statistics [69]. It is also relevant to the classification of the Anderson transitions [77] and topological phases [80, 65] in non-Hermitian systems. For a given non-Hermitian matrix , we introduce a Hermitized matrix
(13) |
Singular values of are given as eigenvalues of and , and hence coincide with singular values of . Thus, the singular-value statistics of non-Hermitian matrices reduce to those of the corresponding Hermitian matrix . We provide the singular-value statistics of Hermitian random matrices in Appendix A (Table 6). Importantly, Hermitized matrices respect additional chiral symmetry by construction,
(14) |
which changes the relevant symmetry classes. For example, for non-Hermitian matrices that respect only time-reversal symmetry in Eq. (1) (i.e., classes AI or AII), the Hermitized matrices belong to classes BDI or CII. We also summarize the correspondence of symmetry classes in Table 1, as well as Tables 2-5.
Through Hermitization, the singular-value statistics of non-Hermitian random matrices with time-reversal symmetry in Eq. (1), particle-hole symmetry in Eq. (2), time-reversal symmetry† in Eq. (4), or particle-hole symmetry† in Eq. (5) reduce to the real-eigenvalue statistics of Hermitized matrices in the chiral [103, 104, 105, *Damgaard-01] and BdG [6, 107] classes. In fact, this correspondence is consistent with our numerical calculations of the singular-value statistics (compare Table 1 and Fig. 1 with Table 6 and Fig. 4, respectively). The joint probability distribution functions of singular values ’s are given as
(15) |
By contrast, in the presence of chiral symmetry in Eq. (6), the Hermitized matrices satisfy
(16) |
and hence the singular values of further reduce to the singular values of the Hermitian matrices and without any symmetry (i.e., class A). When the non-Hermitian matrix is taken from the Gaussian ensemble, the Hermitian matrices and also belong to the Gaussian ensemble. Since singular values of Hermitian matrices are given as the absolute values of their eigenvalues, the singular-value statistics without symmetry reduce to the real-eigenvalue statistics of two independent ensembles one of which corresponds to positive eigenvalues and the other of which corresponds to negative eigenvalues. There, uncorrelated levels can be close to each other, and the level repulsion is significantly weakened, which is the origin of in Fig. 1 (b). Still, a part of the level repulsion survives, resulting in the deviation from the Poisson statistics.
Furthermore, non-Hermitian matrices with sublattice symmetry in Eq. (3) can be generally expressed as
(17) |
with two non-Hermitian matrices and , where we assume . Hence, the singular values of consist of the singular values of the two independent non-Hermitian matrices and . Consequently, the spacing distributions in the bulk of the singular-value spectra again reduce to those of Hermitian matrices without symmetry. On the other hand, the minimum singular value of reduces to the smaller one of the minimum singular values of and , the distribution of which obeys that of Hermitian matrices with chiral symmetry (a subtle difference can arise in generic cases; see Sec. III.3 for details). In the absence of chiral or sublattice symmetry, positive and negative eigenvalues of Hermitized matrices are correlated owing to symmetry, leading to the level repulsion between singular values.
III.3 Classification
Based on Hermitization, we completely classify the singular-value statistics of non-Hermitian random matrices for all the 38 symmetry classes in Tables 2-5, 9 of which are summarized also in Table 1. In a similar manner to the 10-fold AZ symmetry classes for Hermitian matrices, time-reversal symmetry in Eq. (1), particle-hole symmetry in Eq. (2), and chiral symmetry in Eq. (6) form the 10-fold symmetry classes for non-Hermitian matrices (Tables 2 and 3). Furthermore, time-reversal symmetry† in Eq. (4), particle-hole symmetry† in Eq. (5), and chiral symmetry in Eq. (6) form additional 10-fold symmetry classes, which are the AZ† symmetry classes for non-Hermitian matrices (Tables 2 and 4). Taking sublattice symmetry into consideration as additional symmetry (Tables 2 and 5), we have the 38-fold symmetry classification for non-Hermitian matrices [65]. Some symmetry classes in Tables 2-5 give the equivalent symmetry classes, which are not double counted in the 38-fold symmetry classification. For example, time-reversal symmetry in Eq. (1) and particle-hole symmetry† in Eq. (5) lead to essentially the same universal spectral statistics of both eigenvalues and singular values [108]. Consequently, classes AI and AII are equivalent to classes D† and C†, respectively (Table 3).
In these classification tables, we provide the random-matrix indices for the singular-value statistics in the bulk of the spectra [i.e., Eq. (8)] and for the statistics of the minimum singular value [Eq. (11)]. On the other hand, the symmetry classes for which the corresponding Hermitized matrices belong to the standard classes (i.e., classes A, AI, and AII) are not characterized by these random-matrix indices , as discussed above. In the symmetry classes for which the Hermitized matrices respect additional unitary symmetry and hence are block diagonalized, the singular-value statistics in the bulk of the spectra are no longer described by , whereas the statistics of the minimum singular value can still be described by . Furthermore, we analytically derive the singular-value statistics of small random matrices in Appendix B. These small- analytical results well describe the large- results, in the similar spirit to the Wigner surmise [7, *Wigner-58, 9, 97]. This also contrasts with the complex-eigenvalue statistics of non-Hermitian random matrices, where a significant deviation arises between the large- and small- results [43, 58].
In the presence of chiral symmetry in Eq. (6), singular values of a non-Hermitian matrix generally reduce to singular values of Hermitian matrices and with certain symmetry, as shown in Eq. (16). It should be noted that the spectral supports of and can differ from each other for generic random matrices, as well as physical models such as Lindbladians and non-Hermitian Hamiltonians (see Sec. IV below). The two spectra may not overlap when all the eigenvalues of () are positive (negative). In such a case, the level repulsion between eigenvalues is manifested also in the singular-value statistics, which contrasts with the weaker level repulsion discussed in Sec. III.2. Conversely, for the non-Hermitian matrix from the Gaussian ensemble, the two Hermitian matrices and also belong to the Gaussian ensemble and share the same spectral support. Similarly, to obtain the universal results in the presence of chiral symmetry, the spectrum of the Hermitian matrices should be chosen to be statistically symmetric with respect to zero.
Furthermore, in the presence of sublattice symmetry in Eq. (3), singular values of a non-Hermitian matrix generally decompose into singular values of two independent non-Hermitian matrices and in Eq. (17). In particular, the minimum singular value of reduces to the smaller one of the minimum singular values of and . In the sole presence of sublattice symmetry (i.e., class AIII† in Table 2), the distribution of the minimum singular value in coincides with that in or . However, this is not necessarily the case, in general. For example, in the additional presence of time-reversal symmetry that commutes with sublattice symmetry (i.e., class AI + in Table 5), each non-Hermitian matrix or respects time-reversal symmetry and belongs to class AI. According to the analytical result, the distribution of the minimum singular value in or is [see Eq. (51) in Appendix B]
(18) |
and the cumulative distribution function is
(19) |
Then, the cumulative distribution function for the original non-Hermitian matrix is obtained as
(20) |
leading to
(21) |
Still, there is merely a slight difference between and . In fact, both of them do not vanish for and hence are characterized by . Additionally, we have for and for , which are nearly equal to each other. We also numerically obtain from realizations of non-Hermitian random matrices in the Gaussian ensemble, which is consistent with the analytical result .
Class | CS | SLS | Classifying space | Hermitization | ||
---|---|---|---|---|---|---|
A | AIII | |||||
AIII = A + | A | N/A (A) | N/A (A) | |||
AIII | AIII | |||||
A = AIII† | AIII AIII | N/A (A) | ||||
AIII | A A | N/A (A) | N/A (A) |
Class | TRS | PHS | CS | Classifying space | Hermitization | ||
---|---|---|---|---|---|---|---|
AI = D† | +1 | 0 | 0 | BDI | |||
BDI | +1 | +1 | 1 | D | |||
D | 0 | +1 | 0 | DIII | |||
DIII | -1 | +1 | 1 | AII | N/A (AII) | N/A (AII) | |
AII = C† | -1 | 0 | 0 | CII | |||
CII | -1 | -1 | 1 | C | |||
C | 0 | -1 | 0 | CI | |||
CI | +1 | -1 | 1 | AI | N/A (AI) | N/A (AI) |
Class | TRS† | PHS† | CS | Classifying space | Hermitization | ||
---|---|---|---|---|---|---|---|
AI† | +1 | 0 | 0 | CI | |||
BDI† | +1 | +1 | 1 | AI | N/A (AI) | N/A (AI) | |
D† = AI | 0 | +1 | 0 | BDI | |||
DIII† | -1 | +1 | 1 | D | |||
AII† | -1 | 0 | 0 | DIII | |||
CII† | -1 | -1 | 1 | AII | N/A (AII) | N/A (AII) | |
C† = AII | 0 | -1 | 0 | CII | |||
CI† | +1 | -1 | 1 | C |
Class | Classifying space | Hermitization | ||
BDI + | BDI | |||
DIII + = BDI + | DIII | |||
CII + | CII | |||
CI + = CII + | CI | |||
AI + = AII + | AIII | |||
BDI + = DIII + | A | N/A (A) | N/A (A) | |
D + | AIII | |||
DIII + = BDI + | A | N/A (A) | N/A (A) | |
AII + = AI + | AIII | |||
CII + = CI + | A | N/A (A) | N/A (A) | |
C + | AIII | |||
CI + = CII + | A | N/A (A) | N/A (A) | |
BDI + = DIII + | DIII | |||
DIII + | CII | |||
CII + = CI + | CI | |||
CI + | BDI | |||
AI + | BDI BDI | N/A (AI) | ||
BDI + | D D | N/A (A) | ||
D + | DIII DIII | N/A (AII) | ||
DIII + | AII AII | N/A (AII) | N/A (AII) | |
AII + | CII CII | N/A (AII) | ||
CII + | C C | N/A (A) | ||
C + | CI CI | N/A (AI) | ||
CI + | AI AI | N/A (AI) | N/A (AI) |
IV Dissipative quantum chaos
Eigenvalues of nonintegrable quantum systems isolated from the environment obey the statistics of Hermitian random matrices, characterizing their chaotic behavior [4, 2, 3, 1]. As its generalization to dissipative quantum chaos, we show that singular values of nonintegrable open quantum systems obey the statistics of non-Hermitian random matrices. To this end, we study the singular-value statistics of nonintegrable Lindbladians (Sec. IV.1) and non-Hermitian Hamiltonians (Sec. IV.2).

.
IV.1 Lindbladians
We investigate open quantum systems described by the quantum master equation with the Lindbladian [109, 110, 23, 24, 25]
(22) |
where is a Hermitian Hamiltonian for the coherent dynamics, and ’s are dissipators for the coupling with the surrounding environment. Here, we choose the Hamiltonian as the quantum Ising model
(23) |
where the open boundary conditions are imposed, and is randomly chosen from for each site to break unwanted symmetry. We choose the dissipators as damping
(24) |
or dephasing
(25) |
We study the singular-value statistics of the Lindbladian by doubling the Hilbert space. Specifically, we map the density operator to a pure state in the double Hilbert space. Through this vectorization procedure , the Lindblad equation reduces to , where is the non-Hermitian operator in the double Hilbert space:
(26) |
We calculate the spacing ratios of singular values for the Lindbladians with the damping and dephasing (Fig. 2). There, we focus on singular values of the traceless non-Hermitian operator by shifting the Lindbladian by a constant .
Since arbitrary Lindbladians are required to preserve Hermiticity of density operators, they are invariant under modular conjugation (see, for example, Ref. [63]),
(27) |
where is an antiunitary operator that exchanges the bra and ket degrees of freedom,
(28) |
for arbitrary bosonic operators and complex numbers . Consequently, singular values of the damped Ising model obey the random-matrix statistics in class AI [Fig. 2 (a)]. As also discussed above, time-reversal symmetry affects complex eigenvalues only around the real axis of the Lindbladian spectrum; conversely, time-reversal symmetry even changes the local correlations of arbitrary singular values and manifests itself more clearly in the singular-value statistics.
On the other hand, the dephased Ising model additionally respects time-reversal symmetry†
(29) |
leading to the singular-value statistics in class BDI† [Fig. 2 (b)]. There, chiral symmetry arising from the combination of Eqs. (27) and (29) weakens the level repulsion between singular values even in the chaotic regime. We emphasize that the Lindbladians are represented by sparse matrices because of the locality constraint, as opposed to random matrices. Nevertheless, the singular-value statistics of follow the random-matrix statistics, providing a defining feature of chaos and nonintegrability in open quantum systems.
IV.2 Non-Hermitian Hamiltonians

In addition to the Lindbladians, we also show that singular values of nonintegrable non-Hermitian Hamiltonians follow the random-matrix statistics in the corresponding symmetry classes. We study a non-Hermitian spin model [111, 43]
(30) |
under the open boundary conditions. Here, is randomly chosen from for each site to break unwanted symmetry (e.g., translation symmetry), and is the degree of the Dzyaloshinskii-Moriya interaction. Owing to the nonintegrability of this non-Hermitian spin model , singular values obey the random-matrix statistics (Fig. 3). Depending on the parameters, belongs to classes A, AI†, and AII†, as follows:
-
(i)
For and , the non-Hermitian spin model does not respect any symmetries and belongs to class A [Fig. 3 (a)].
-
(ii)
For and , the non-Hermitian spin model respects time-reversal symmetry†
(31) and belongs to class AI† [Fig. 3 (b)].
-
(iii)
For and , the non-Hermitian spin model respects time-reversal symmetry†
(32) and belongs to class AI† for even and class AII† for odd [Fig. 3 (c)]. For odd , all the singular values are two-fold degenerate.
Time-reversal symmetry† also changes the universal level-spacing statistics for complex eigenvalues [43]. Nevertheless, it does not impact the level repulsion for small level spacing but rather influences the peak of the probability distribution. Conversely, time-reversal symmetry† even affects the level repulsion of singular values, as illustrated in Fig. 3. The distinctive singular-value statistics among the different symmetry classes are clearly discerned according to the power-law behavior for small level spacing, as in Eq. (8).
V Discussions
The spectral statistics play a pivotal role in elucidating the nature of chaos and integrability in both closed and open quantum systems, forming the fundamental basis of statistical mechanics. In this work, we study the statistical properties of singular values, as opposed to complex eigenvalues, in non-Hermitian random matrices, offering a powerful diagnostic tool for characterizing dissipative quantum chaos. By means of Hermitization, we establish the exhaustive classification of the singular-value statistics for all the 38-fold symmetry classes of non-Hermitian random matrices, summarized as Tables 2-5. We also analytically obtain the singular-value statistics of small random matrices, which well describe those of small random matrices in the similar spirit to the Wigner surmise. Furthermore, we demonstrate that singular values of nonintegrable Lindbladians and non-Hermitian Hamiltonians conform to the random-matrix statistics, enabling the identification of chaos and nonintegrability in open quantum many-body systems.
Notably, we show that the singular-value statistics provide a useful measure that quantifies dissipative quantum chaos, providing distinct information compared to the complex-eigenvalue statistics. Specifically, the correlations of complex eigenvalues in the bulk of the spectra are governed solely by time-reversal symmetry† in Eq. (4) [43], while the other symmetries play a role only in the vicinity of the symmetric lines or points [58], which makes the role of symmetry in dissipative quantum chaos elusive. By contrast, symmetries irrelevant to the complex-eigenvalue statistics in the bulk, such as time-reversal symmetry in Eq. (1), change the spectral statistics of singular values even in the bulk. Additionally, the difference in the bulk complex-eigenvalue statistics among the symmetry classes is much more subtle compared to Hermitian random matrices [43]. For example, the level repulsion of complex eigenvalues is universally cubic regardless of time-reversal symmetry† [37]. Conversely, we show that the level repulsion of singular values relies on symmetry as in Eq. (8), reminiscent of the Wigner-Dyson universality of Hermitian random matrices. Thus, the singular-value statistics yield a clearer indicator of symmetry in open quantum systems. The combination of the two types of spectral statistics—complex-eigenvalue and singular-value statistics—leads to a more profound understanding of dissipative quantum chaos.
One of the significant applications of the spectral statistics is the identification of phase transitions between the chaotic and integrable regimes, including the Anderson transitions. In a similar manner to the eigenvalue statistics, the singular-value statistics should be utilized to obtain the universal critical exponents and scaling functions of the Anderson transitions, which we leave for future work. In this respect, it is noteworthy that the critical points of the Anderson transitions generally depend on the eigenvalues. Consequently, to capture the critical behavior precisely, we need to focus on the spectral statistics within a fixed window of eigenvalues. However, for non-Hermitian systems, the complex-eigenvalue window includes fewer eigenstates since complex eigenvalues are distributed in the two-dimensional complex plane, thereby making the numerical finite-size scaling more challenging [75]. On the other hand, owing to the real-valued nature of the singular values, even in non-Hermitian systems, qualitatively more singular values are included in a fixed window of singular values. Thus, the singular-value statistics should be advantageous for precisely characterizing the Anderson transitions in open systems.
Finally, it is also worthwhile to further explore the fundamental role of singular values in the physics of open quantum systems. The singular-value statistics are also relevant to closed quantum systems. As a prime example, reflection matrices in the scattering process are generally non-Hermitian, even when the corresponding Hamiltonian is Hermitian. Accordingly, singular values of reflection matrices, square roots of the reflection probability, describe various quantum transport phenomena, such as mesoscopic electronic transport and nuclear reactions. Notably, for Hermitian Hamiltonians within the AZ symmetry class, the corresponding non-Hermitian reflection matrices generally belong to the AZ† symmetry class, as demonstrated in Appendix D. Thus, the statistical properties of singular values in non-Hermitian reflection matrices are described by our classification tables, especially those for the 10-fold AZ† symmetry class in Tables 2 and 4.
Acknowledgements.
K.K. is supported by the Japan Society for the Promotion of Science (JSPS) through the Overseas Research Fellowship, and by the Gordon and Betty Moore Foundation through Grant No. GBMF8685 toward the Princeton theory program. Z.X. and R.S. are supported by the National Basic Research Programs of China (No. 2019YFA0308401) and by the National Natural Science Foundation of China (No. 11674011 and No. 12074008). T.O. is supported by JSPS KAKENHI Grants No. 19H00658 and No. 22H05114, and CREST Grant No. JPMJCR18T4.AZ class | TRS | PHS | CS | ||||||
A | N/A | N/A | |||||||
AIII | |||||||||
AI | N/A | N/A | |||||||
BDI | |||||||||
D | |||||||||
DIII | |||||||||
AII | N/A | N/A | |||||||
CII | |||||||||
C | |||||||||
CI |

Appendix A Singular-value statistics of Hermitian random matrices
As discussed in Sec. III.2, the singular-value statistics of non-Hermitian random matrices reduce to those of Hermitian random matrices. Here, we study the singular-value statistics of Hermitian random matrices in all the 10-fold AZ symmetry classes (Table 6 and Fig. 4). Figure 4 (a-c) shows the density of singular values whose spectral radius is normalized to be unity. Away from zero singular value , it is described by the Wigner semicircle law
(33) |
for all the symmetry classes. Equivalently, squared singular values follow the Marchenko-Pastur distribution [112]
(34) |
for square matrices. Around , on the other hand, the density of singular values exhibits different behaviors depending on symmetry classes, as discussed below.
Figure 4 (d-f) shows the distributions of the level-spacing ratios of singular values in Eq. (7). The average of gives a useful measure that distinguishes between different symmetry classes, as summarized in Table 6. The singular-value spectra exhibit the two-fold (four-fold) degeneracy in classes AIII, BDI, D, AII, C, and CI (DIII and CII). In the chiral and BdG classes, eigenvalues come in opposite-sign pairs because of chiral or particle-hole symmetry, and hence the level-spacing-ratio (or level-spacing) distributions of singular values reduce to those of eigenvalues [96, 97], which are determined only by time-reversal symmetry and described by the Dyson index . In the standard classes, by contrast, ’s are not described by any of the conventional eigenvalue statistics. For example, the probability density for zero level-spacing ratio, , is nonvanishing in the standard classes in contrast to the vanishing probability in the chiral and BdG classes. This behavior originates from the fact that uncorrelated levels can be close to each other in the singular-value spectrum without chiral or particle-hole symmetry. In fact, ’s in the standard classes reduce to those of eigenvalues for two independent random matrices. Notably, a part of the spectral correlations still survives, which leads to the 3-fold distributions with but for the Poisson distribution in Eq. (10). In Appendix B, we analytically derive the level-spacing-ratio distributions of singular values for small random matrices. These analytical results well describe the level-spacing-ratio distributions for large random matrices in the similar spirit to the Wigner surmise.
Figure 4 (g-i) shows the distributions of the minimum singular value whose average is normalized to be unity (i.e., ). Here, the average normalized variance of the minimum singular value, , gives a useful measure that distinguishes between different symmetry classes and does not depend on the normalization of singular values. In the chiral and BdG classes, the distributions of the minimum singular value, , reduce to the distributions of the minimum nonnegative eigenvalue. In the standard classes, by contrast, ’s do not reduce to any of the eigenvalue statistics. In Appendix B, we analytically derive for small random matrices in the standard classes. Importantly, the combination of the level-spacing-ratio (or level-spacing) distributions and the distributions of the minimum singular value completely distinguishes the 10-fold symmetry classes, each of which is described by the Dyson index in Eq. (8) and the other random-matrix index in Eq. (11), respectively. While some symmetry classes can exhibit the same level-spacing-ratio distribution (e.g., classes AIII, D, and C), the distributions of the minimum singular value are different between different symmetry classes.
Appendix B Wigner surmise of singular-value statistics
We analytically derive the singular-value statistics of small Hermitian random matrices in all the 10 symmetry classes. Wigner proposed using small Hermitian random matrices to capture the level-spacing distribution of sufficiently complex nuclei of heavy atoms [7, *Wigner-58]. Similarly, we show that the analytical results of the singular-value statistics for small random matrices well capture those for large random matrices and also physical systems in the chaotic regime. As also discussed in Appendix A, the singular-value statistics in the chiral and BdG classes reduce to the eigenvalue statistics. In the standard (Wigner-Dyson) classes, by contrast, the singular-value statistics do not reduce to any of the conventional eigenvalue statistics and exhibit the weaker level repulsion than those in the chiral and BdG classes (Fig. 5). In Table 6, we summarize the average level-spacing ratios and the average normalized variances of the minimum singular value for small random matrices, which well describe the singular-value statistics of large random matrices and serve as useful measures that distinguish between different symmetry classes.

B.1 Level-spacing distribution
In the chiral and BdG classes, the level-spacing distributions of singular values reduce to the level-spacing distributions of nonnegative eigenvalues, leading to [100, 1]
(35) |
where is normalized by . We have
(36) |
In the standard classes, the smallest random matrices to study the level-spacing distributions of singular values are () Hermitian random matrices for classes A and AI (class AII). The joint probability distribution function of the two eigenvalues in the Gaussian ensemble is given as
(37) |
where we assume , and the normalization constant is
(38) |
The distributions of the level spacing between the two singular values, , are obtained as
(39) |
We normalize by with
(40) |
which satisfies . We also have
(41) |
Notably, the probability distribution functions do not vanish even for small level spacing in contrast with the chiral and BdG classes. This is because the two singular values can be close to each other in the absence of chiral and particle-hole symmetry even if the corresponding two eigenvalues are away from each other. Still, a part of the level repulsion survives, which leads to and with the Poisson statistics .
B.2 Level-spacing-ratio distribution
In the chiral and BdG classes, the level-spacing-ratio distributions of singular values reduce to the level-spacing-ratio distributions of nonnegative eigenvalues in the chiral and BdG classes, similarly to the level-spacing distributions, leading to [97]
(42) |
with the normalization constants
(43) |
The average level-spacing ratio is obtained as
(44) |
In the standard classes, the smallest random matrices to study the level-spacing-ratio distributions of singular values are () Hermitian random matrices for classes A and AI (class AII). The joint probability distribution function of the three eigenvalues in the Gaussian ensemble is
(45) |
where we assume and , and the normalization constant is
(46) |
We have four possible cases: (i) ; (ii) ; (iii) ; (iv) . For example, for the case (i), we have three singular values . Let us introduce and , which satisfy , , and . The level-spacing ratio of the singular values is for and for . Then, the level-spacing-ratio distribution of the singular values for the case (i) is obtained as
(47) |
The other three cases can be evaluated in a similar manner, and the sum of the four terms is given by
(48) |
For arbitrary , this integral can be expressed by elementary but complicated functions. We have
(49) |
and
(50) |
While the level repulsion arises even in the singular-value spectrum for the cases (i), (iii), and (iv), it is weakened for the case (ii), which results in the nonzero probability density for .
B.3 Distribution of the minimum singular value
In the presence of chiral or particle-hole symmetry, the distributions of the minimum singular value reduce to the distributions of the minimum nonnegative eigenvalue. In the chiral classes, they are exactly obtained as [103, 104, 105, *Damgaard-01]
(51) |
where is the modified Bessel function of the first kind. We then have
(52) |
In the BdG classes, we have [107]
(53) |
and
(54) |
In the standard classes, the smallest random matrices to study the distribution of the minimum singular value are () Hermitian random matrices for classes A and AI (class AII). The joint probability distribution function of the two eigenvalues in the Gaussian ensemble is given as Eq. (37), where we assume . The distributions of the minimum singular value are obtained as
(55) |
We normalize by with
(56) |
which satisfies . We also have
(57) |
As shown in Fig. 5, while the small- results well describe the large- results for classes A and AI, a significant deviation is observed for class AII.
To explain the deviation of the small- result from the large- result in class AII, we further calculate the distribution of the minimum singular value for Hermitian random matrices in class AII with larger sizes (). For a generic number of eigenvalues, the joint probability density function of eigenvalues , , , in the Gaussian ensemble is
(58) |
with the normalization constant . We examine the probability of all singular values being larger than , given as
(59) |
From , the distribution of the minimum singular value is obtained as
(60) |
Expanding the term in Eq. (58) as a polynomial, we find that each term takes the form,
(61) |
Notably, after performing the integral in Eq. (59), only the terms with all , , being even contribute, and the other terms vanish. Meanwhile, since ’s are only dummy variables, we can rewrite as , where , , , represent the values arranged in descending order. The integral can be decomposed into independent Gaussian integrals, each of which is given as
(62) |
with the incomplete gamma function . Thus, in principle, we can calculate for arbitrary . This calculation can be straightforwardly carried out by a computer algebra system. For example, with is obtained as
(63) |

In this manner, we analytically calculate for (i.e., , , Hermitian random matrices in class AII) [Fig. 6 (a)]. Notably, we find a significant even-odd effect depending on the number of different eigenvalues. Such an even-odd effect disappears for large random matrices [Fig. 6 (b)], and converges to the large- results in Fig. 5 (i) that lie between the small-even and small-odd results.
To understand this even-odd effect, let us consider the eigenvalues , , , arranged in ascending order. Since the eigenvalue spectrum statistically distributes in a symmetric manner with respect to zero, one can expect that the smallest singular value corresponds to the eigenvalue near the middle of the spectrum. Then, for even , the two eigenvalues and can be the smallest singular value, and the situation for and is most likely. Without loss of generality, we can assume that contributes to the smallest singular value. In this scenario, the eigenvalue experiences the asymmetric level repulsion due to the presence of negative eigenvalues and positive eigenvalues. This asymmetry of the level repulsion becomes significant for small . The stronger level repulsion due to the negative eigenvalues prevents the eigenvalue from approaching zero closely, which causes the peak of to shift away from the origin [see Fig. 6 (a)]. For odd , on the other hand, lies in the middle of the eigenvalue spectrum and should contribute to the smallest singular value. The most likely situation is that half of the remaining eigenvalues are positive while the other half are negative. In this scenario, the middle eigenvalue experiences the equal strength of level repulsion from both the negative and positive eigenvalues. Consequently, it tends to approach zero in the eigenvalue spectrum, leading to the peak at the origin in the distribution . In contrast to class AII, the even-odd effect should not be significant even for small random matrices in classes A and AI owing to the weaker level repulsion.
Appendix C Normal random matrices
Normal matrices are an important family of matrices. A normal matrix is required to satisfy
(64) |
as its defining property. Hermitian matrices are a subfamily of normal matrices. While generic non-Hermitian matrices are not necessarily diagonalized by unitary matrices, a normal matrix is always diagonalized by a unitary matrix ,
(65) |
where is a diagonal matrix composed of complex eigenvalues . One way of generating an ensemble of normal matrices is to start from an ensemble of generic random matrices and selectively include normal matrices while excluding nonnormal ones. A numerical algorithm for generating such an ensemble of normal random matrices was provided, for example, in Ref. [113]. The joint probability distribution function of eigenvalues for normal random matrices without any symmetry (i.e., class A) in the Gaussian ensemble was shown to be identical to that of generic non-Hermitian random matrices [113, 114, 115]. Consequently, generic and normal random matrices share the same eigenvalue-spacing and eigenvalue-spacing-ratio statistics.
In contrast with the complex-eigenvalue statistics, the singular-value statistics exhibit distinctions between generic and normal random matrices. While singular values and eigenvalues are not related to each other in generic non-Hermitian matrices, singular values ’s of a normal matrix are given as the absolute values of its eigenvalues,
(66) |
Then, since the eigenvalues ’s of a normal matrix are distributed uniformly in a circle in the complex plane, the density of the singular values ’s is given as
(67) |
with the spectral radius normalized to unity. Remarkably, this clearly deviates from the semicircle law in Eq. (33), which is respected for generic non-Hermitian random matrices.
Furthermore, spacings and spacing ratios of singular values follow the Poisson statistics for normal random matrices. For two adjacent singular values, and , the corresponding eigenvalues, and , can be well separated from each other due to different phases, and . Specifically, the probability that and are adjacent complex eigenvalues, which entails with being a constant on the order of the mean level spacing of complex eigenvalues, is proportional to . Given that is of the order for the matrix size , the probability scales as and vanishes entirely as , suggesting the vanishing repulsion between the adjacent singular values. Notably, this situation differs from the Hermitian case, where the repulsion between singular values similarly weakens but a part of it persists. This difference arises because Hermiticity makes eigenvalues ’s real and enforces or . In the presence of such a constraint, the probability of scales as , leading to the deviations from the Poisson statistics.
It is also worthwhile to study the singular-value statistics of normal random matrices in the presence of symmetry. As an illustrative example, we here study real normal random matrices (i.e., class AI) and show that time-reversal symmetry has a significant impact on the singular-value statistics. A generic real matrix can be parameterized as
(68) |
with and Pauli matrices (). The normality condition in Eq. (64) requires the constraints and , further leading to (i) and , or (ii) . Within the ensemble of normal random matrices induced from the larger ensemble, the probability of the case (ii) is much higher than that of the case (i) since fine-tuning of the additional parameter is necessary for the case (i). Thus, in the ensemble of normal random matrices with time-reversal symmetry, almost all matrices are required to respect Hermiticity, resulting in the distinct statistics of eigenvalues and singular values. We leave more systematic exploration on the interplay of normality and symmetry for future study.
Appendix D Symmetry classification of non-Hermitian reflection matrices
The singular-value statistics are also relevant to the physics of closed quantum systems. As a prime example, reflection matrices in the scattering process are generally non-Hermitian, even when the corresponding Hamiltonian is Hermitian. Singular values of these reflection matrices, square roots of the reflection probability, describe quantum transport phenomena. The fluctuations of singular values from sample to sample, or the suppression thereof, are reflected in the conductance fluctuations [89, *Lee-87, 91, 92, 93]. Here, we show that if original Hermitian Hamiltonians belong to the AZ symmetry class, the corresponding non-Hermitian reflection matrices generally belong to the AZ† symmetry class. Specifically, we establish that time-reversal symmetry in Eq. (1) and particle-hole symmetry in Eq. (2) of Hamiltonians impose time-reversal symmetry† in Eq. (4) and particle-hole symmetry† in Eq. (5) on reflection matrices. Similarly, chiral symmetry in Eq. (3) of Hamiltonians leads to chiral symmetry in Eq. (6) of reflection matrices. Thus, the singular-value statistics of non-Hermitian reflection matrices are described by our classification tables for the 10-fold AZ† symmetry class (Tables 2 and 4).
The following discussions are based on the Mahaux-Weidenmüller formula [116, *Mahaux-Weidenmuller-69, 118, 14, *Beenakker-review-15],
(69) |
where is a single-particle energy of the incident and scattered waves, and is a coupling matrix between the system and the ideal leads that is assumed to share the same symmetries as the system. The scattering matrix reads [87, 88],
(70) |
where () is a reflection matrix from the left to the left (from the right to the right), and () is a transmission matrix from the left to the right (from the right to the left). If the Hamiltonian respects Hermiticity (), the scattering matrix respects unitarity (). We stress that the reflection matrix is generally non-Hermitian even if the corresponding Hamiltonian is Hermitian. In Ref. [119, *Zirnstein-21B], the Mahaux-Weidenmüller formula is used to understand the bulk-boundary correspondence of non-Hermitian Hamiltonians. Here, we instead consider Hermitian Hamiltonians and study symmetry of the corresponding non-Hermitian reflection matrices. When Hamiltonians lack Hermiticity, reflection matrices are no longer described solely by the AZ† symmetry class but are still categorized into one of the 38 symmetry classes.
D.1 Time-reversal symmetry
Suppose that Hermitian Hamiltonians respect time-reversal symmetry in Eq. (1), leading to
(71) |
and
(72) |
While typically satisfies for ideal leads, we only need with arbitrary to have these relations (this is also the case for particle-hole and chiral transformations below). Thus, we have
(73) |
which means that the reflection matrix respects time-reversal symmetry† in Eq. (4) for arbitrary . It should be noted that the matrix sizes of are different between Eqs. (72) and (73). While in Eq. (72) acts on the incoming and outgoing modes on both left and right sides, in Eq. (73) acts solely on those on either left or right side.
D.2 Particle-hole symmetry
Suppose that Hamiltonians respect particle-hole symmetry in Eq. (2), leading to
(74) |
and
(75) |
Thus, for zero modes , we have
(76) |
which means that the reflection matrix respects time-reversal symmetry, or equivalently, respects particle-hole symmetry† in Eq. (5). It should be noted that particle-hole symmetry† is respected only for zero modes and is explicitly broken for .
D.3 Chiral symmetry
References
- Haake et al. [2018] F. Haake, S. Gnutzmann, and M. Kuś, Quantum Signatures of Chaos (Springer, Cham, 2018).
- Nandkishore and Huse [2015] R. Nandkishore and D. A. Huse, Many-Body Localization and Thermalization in Quantum Statistical Mechanics, Annu. Rev. Condens. Matter Phys. 6, 15 (2015).
- D’Alessio et al. [2016] L. D’Alessio, Y. Kafri, A. Polkovnikov, and M. Rigol, From quantum chaos and eigenstate thermalization to statistical mechanics and thermodynamics, Adv. Phys. 65, 239 (2016).
- Bohigas et al. [1984] O. Bohigas, M. J. Giannoni, and C. Schmit, Characterization of Chaotic Quantum Spectra and Universality of Level Fluctuation Laws, Phys. Rev. Lett. 52, 1 (1984).
- Berry and Tabor [1977] M. V. Berry and M. Tabor, Level clustering in the regular spectrum, Proc. R. Soc. A 356, 375 (1977).
- Altland and Zirnbauer [1997] A. Altland and M. R. Zirnbauer, Nonstandard symmetry classes in mesoscopic normal-superconducting hybrid structures, Phys. Rev. B 55, 1142 (1997).
- Wigner [1951] E. P. Wigner, On the statistical distribution of the widths and spacings of nuclear resonance levels, Math. Proc. Cambridge Philos. Soc. 47, 790 (1951).
- Wigner [1958] E. P. Wigner, On the Distribution of the Roots of Certain Symmetric Matrices, Ann. Math. 67, 325 (1958).
- Dyson [1962] F. J. Dyson, The Threefold Way. Algebraic Structure of Symmetry Groups and Ensembles in Quantum Mechanics, J. Math. Phys. 3, 1199 (1962).
- Gade and Wegner [1991] R. Gade and F. Wegner, The replica limit of and models, Nucl. Phys. 360, 213 (1991).
- Gade [1993] R. Gade, Anderson localization for sublattice models, Nucl. Phys. 398, 499 (1993).
- Verbaarschot [1994] J. Verbaarschot, Spectrum of the QCD Dirac operator and chiral random matrix theory, Phys. Rev. Lett. 72, 2531 (1994).
- Verbaarschot and Wettig [2000] J. J. M. Verbaarschot and T. Wettig, Random Matrix Theory and Chiral Symmetry in QCD, Annu. Rev. Nucl. Part. Sci. 50, 343 (2000).
- Beenakker [1997] C. W. J. Beenakker, Random-matrix theory of quantum transport, Rev. Mod. Phys. 69, 731 (1997).
- Beenakker [2015] C. W. J. Beenakker, Random-matrix theory of Majorana fermions and topological superconductors, Rev. Mod. Phys. 87, 1037 (2015).
- Evers and Mirlin [2008] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008).
- Schnyder et al. [2008] A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Ludwig, Classification of topological insulators and superconductors in three spatial dimensions, Phys. Rev. B 78, 195125 (2008).
- Ryu et al. [2010] S. Ryu, A. P. Schnyder, A. Furusaki, and A. W. W. Ludwig, Topological insulators and superconductors: tenfold way and dimensional hierarchy, New J. Phys. 12, 065010 (2010).
- Kitaev [2009] A. Kitaev, Periodic table for topological insulators and superconductors, AIP Conf. Proc. 1134, 22 (2009).
- Hasan and Kane [2010] M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).
- Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
- Chiu et al. [2016] C.-K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, Classification of topological quantum matter with symmetries, Rev. Mod. Phys. 88, 035005 (2016).
- Nielsen and Chuang [2000] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, England, 2000).
- Breuer and Petruccione [2007] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, Oxford, 2007).
- Rivas and Huelga [2012] A. Rivas and S. F. Huelga, Open Quantum Systems (Springer, Berlin, Heidelberg, 2012).
- Carmichael [1993] H. Carmichael, An Open Systems Approach to Quantum Optics (Springer, Berlin, 1993).
- Plenio and Knight [1998] M. B. Plenio and P. L. Knight, The quantum-jump approach to dissipative dynamics in quantum optics, Rev. Mod. Phys. 70, 101 (1998).
- Daley [2014] A. J. Daley, Quantum trajectories and open many-body quantum systems, Adv. Phys. 63, 77 (2014).
- Bender [2007] C. M. Bender, Making sense of non-Hermitian Hamiltonians, Rep. Prog. Phys. 70, 947 (2007).
- Konotop et al. [2016] V. V. Konotop, J. Yang, and D. A. Zezyulin, Nonlinear waves in -symmetric systems, Rev. Mod. Phys. 88, 035002 (2016).
- El-Ganainy et al. [2018] R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, Non-Hermitian physics and PT symmetry, Nat. Phys. 14, 11 (2018).
- Potter and Vasseur [2022] A. C. Potter and R. Vasseur, Entanglement Dynamics in Hybrid Quantum Circuits, in Entanglement in Spin Chains: From Theory to Quantum Technology Applications, edited by A. Bayat, S. Bose, and H. Johannesson (Springer, Cham, 2022) pp. 211–249.
- Fisher et al. [2023] M. P. A. Fisher, V. Khemani, A. Nahum, and S. Vijay, Random Quantum Circuits, Annu. Rev. Condens. Matter Phys. 14, 335 (2023).
- Bergholtz et al. [2021] E. J. Bergholtz, J. C. Budich, and F. K. Kunst, Exceptional topology of non-Hermitian systems, Rev. Mod. Phys. 93, 015005 (2021).
- Okuma and Sato [2023] N. Okuma and M. Sato, Non-Hermitian Topological Phenomena: A Review, Annu. Rev. Condens. Matter Phys. 14, 83 (2023).
- Grobe et al. [1988] R. Grobe, F. Haake, and H.-J. Sommers, Quantum Distinction of Regular and Chaotic Dissipative Motion, Phys. Rev. Lett. 61, 1899 (1988).
- Grobe and Haake [1989] R. Grobe and F. Haake, Universality of cubic-level repulsion for dissipative quantum chaos, Phys. Rev. Lett. 62, 2893 (1989).
- Xu et al. [2019] Z. Xu, L. P. García-Pintos, A. Chenu, and A. del Campo, Extreme Decoherence and Quantum Chaos, Phys. Rev. Lett. 122, 014103 (2019).
- Hamazaki et al. [2019] R. Hamazaki, K. Kawabata, and M. Ueda, Non-Hermitian Many-Body Localization, Phys. Rev. Lett. 123, 090603 (2019).
- Denisov et al. [2019] S. Denisov, T. Laptyeva, W. Tarnowski, D. Chruściński, and K. Życzkowski, Universal Spectra of Random Lindblad Operators, Phys. Rev. Lett. 123, 140403 (2019).
- Can et al. [2019] T. Can, V. Oganesyan, D. Orgad, and S. Gopalakrishnan, Spectral Gaps and Midgap States in Random Quantum Master Equations, Phys. Rev. Lett. 123, 234103 (2019).
- Can [2019] T. Can, Random Lindblad dynamics, J. Phys. A 52, 485302 (2019).
- Hamazaki et al. [2020] R. Hamazaki, K. Kawabata, N. Kura, and M. Ueda, Universality classes of non-Hermitian random matrices, Phys. Rev. Research 2, 023286 (2020).
- Akemann et al. [2019] G. Akemann, M. Kieburg, A. Mielke, and T. Prosen, Universal Signature from Integrability to Chaos in Dissipative Open Quantum Systems, Phys. Rev. Lett. 123, 254101 (2019).
- Sá et al. [2020] L. Sá, P. Ribeiro, and T. Prosen, Complex Spacing Ratios: A Signature of Dissipative Quantum Chaos, Phys. Rev. X 10, 021019 (2020).
- Wang et al. [2020] K. Wang, F. Piazza, and D. J. Luitz, Hierarchy of Relaxation Timescales in Local Random Liouvillians, Phys. Rev. Lett. 124, 100604 (2020).
- Xu et al. [2021] Z. Xu, A. Chenu, T. Prosen, and A. del Campo, Thermofield dynamics: Quantum chaos versus decoherence, Phys. Rev. B 103, 064309 (2021).
- García-García et al. [2022a] A. M. García-García, Y. Jia, D. Rosa, and J. J. M. Verbaarschot, Dominance of Replica Off-Diagonal Configurations and Phase Transitions in a Symmetric Sachdev-Ye-Kitaev Model, Phys. Rev. Lett. 128, 081601 (2022a).
- García-García et al. [2022b] A. M. García-García, Y. Jia, D. Rosa, and J. J. M. Verbaarschot, Replica symmetry breaking in random non-Hermitian systems, Phys. Rev. D 105, 126027 (2022b).
- Li et al. [2021] J. Li, T. Prosen, and A. Chan, Spectral Statistics of Non-Hermitian Matrices and Dissipative Quantum Chaos, Phys. Rev. Lett. 127, 170602 (2021).
- Tarnowski et al. [2021] W. Tarnowski, I. Yusipov, T. Laptyeva, S. Denisov, D. Chruściński, and K. Życzkowski, Random generators of Markovian evolution: A quantum-classical transition by superdecoherence, Phys. Rev. E 104, 034118 (2021).
- Cornelius et al. [2022] J. Cornelius, Z. Xu, A. Saxena, A. Chenu, and A. del Campo, Spectral Filtering Induced by Non-Hermitian Evolution with Balanced Gain and Loss: Enhancing Quantum Chaos, Phys. Rev. Lett. 128, 190402 (2022).
- García-García et al. [2022c] A. M. García-García, L. Sá, and J. J. M. Verbaarschot, Symmetry Classification and Universality in Non-Hermitian Many-Body Quantum Chaos by the Sachdev-Ye-Kitaev Model, Phys. Rev. X 12, 021040 (2022c).
- Sá et al. [2022] L. Sá, P. Ribeiro, and T. Prosen, Lindbladian dissipation of strongly-correlated quantum matter, Phys. Rev. Research 4, L022068 (2022).
- Kulkarni et al. [2022] A. Kulkarni, T. Numasawa, and S. Ryu, Lindbladian dynamics of the Sachdev-Ye-Kitaev model, Phys. Rev. B 106, 075138 (2022).
- Cipolloni and Kudler-Flam [2023a] G. Cipolloni and J. Kudler-Flam, Entanglement Entropy of Non-Hermitian Eigenstates and the Ginibre Ensemble, Phys. Rev. Lett. 130, 010401 (2023a).
- Cipolloni and Kudler-Flam [2023b] G. Cipolloni and J. Kudler-Flam, Non-Hermitian Hamiltonians Violate the Eigenstate Thermalization Hypothesis, arXiv:2303.03448 (2023b).
- Xiao et al. [2022] Z. Xiao, K. Kawabata, X. Luo, T. Ohtsuki, and R. Shindou, Level statistics of real eigenvalues in non-Hermitian systems, Phys. Rev. Research 4, 043196 (2022).
- Shivam et al. [2023] S. Shivam, A. De Luca, D. A. Huse, and A. Chan, Many-Body Quantum Chaos and Emergence of Ginibre Ensemble, Phys. Rev. Lett. 130, 140403 (2023).
- Ghosh et al. [2022] S. Ghosh, S. Gupta, and M. Kulkarni, Spectral properties of disordered interacting non-Hermitian systems, Phys. Rev. B 106, 134202 (2022).
- Matsoukas-Roubeas et al. [2023] A. S. Matsoukas-Roubeas, F. Roccati, J. Cornelius, Z. Xu, A. Chenu, and A. del Campo, Non-Hermitian Hamiltonian deformations in quantum mechanics, J. High Energ. Phys. 2023 (1), 60.
- Sá et al. [2023] L. Sá, P. Ribeiro, and T. Prosen, Symmetry Classification of Many-Body Lindbladians: Tenfold Way and Beyond, Phys. Rev. X 13, 031019 (2023).
- Kawabata et al. [2023] K. Kawabata, A. Kulkarni, J. Li, T. Numasawa, and S. Ryu, Symmetry of Open Quantum Systems: Classification of Dissipative Quantum Chaos, PRX Quantum 4, 030328 (2023).
- Bernard and LeClair [2002] D. Bernard and A. LeClair, A Classification of Non-Hermitian Random Matrices, in Statistical Field Theories, edited by A. Cappelli and G. Mussardo (Springer, Dordrecht, 2002) pp. 207–214.
- Kawabata et al. [2019a] K. Kawabata, K. Shiozaki, M. Ueda, and M. Sato, Symmetry and Topology in Non-Hermitian Physics, Phys. Rev. X 9, 041015 (2019a).
- Hatano and Nelson [1996] N. Hatano and D. R. Nelson, Localization Transitions in Non-Hermitian Quantum Mechanics, Phys. Rev. Lett. 77, 570 (1996).
- Hatano and Nelson [1997] N. Hatano and D. R. Nelson, Vortex pinning and non-Hermitian quantum mechanics, Phys. Rev. B 56, 8651 (1997).
- Efetov [1997] K. B. Efetov, Directed Quantum Chaos, Phys. Rev. Lett. 79, 491 (1997).
- Feinberg and Zee [1997] J. Feinberg and A. Zee, Non-hermitian random matrix theory: Method of hermitian reduction, Nucl. Phys. B 504, 579 (1997).
- Brouwer et al. [1997] P. W. Brouwer, P. G. Silvestrov, and C. W. J. Beenakker, Theory of directed localization in one dimension, Phys. Rev. B 56, R4333(R) (1997).
- Longhi [2019] S. Longhi, Topological Phase Transition in non-Hermitian Quasicrystals, Phys. Rev. Lett. 122, 237601 (2019).
- Tzortzakakis et al. [2020] A. F. Tzortzakakis, K. G. Makris, and E. N. Economou, Non-Hermitian disorder in two-dimensional optical lattices, Phys. Rev. B 101, 014202 (2020).
- Huang and Shklovskii [2020] Y. Huang and B. I. Shklovskii, Anderson transition in three-dimensional systems with non-Hermitian disorder, Phys. Rev. B 101, 014204 (2020).
- Kawabata and Ryu [2021] K. Kawabata and S. Ryu, Nonunitary Scaling Theory of Non-Hermitian Localization, Phys. Rev. Lett. 126, 166801 (2021).
- Luo et al. [2021a] X. Luo, T. Ohtsuki, and R. Shindou, Universality Classes of the Anderson Transitions Driven by Non-Hermitian Disorder, Phys. Rev. Lett. 126, 090402 (2021a).
- Luo et al. [2021b] X. Luo, T. Ohtsuki, and R. Shindou, Transfer matrix study of the Anderson transition in non-Hermitian systems, Phys. Rev. B 104, 104203 (2021b).
- Luo et al. [2022] X. Luo, Z. Xiao, K. Kawabata, T. Ohtsuki, and R. Shindou, Unifying the Anderson transitions in Hermitian and non-Hermitian systems, Phys. Rev. Research 4, L022035 (2022).
- Liu et al. [2021] H. Liu, J.-S. You, S. Ryu, and I. C. Fulga, Supermetal-insulator transition in a non-Hermitian network model, Phys. Rev. B 104, 155412 (2021).
- Ghosh et al. [2023] S. Ghosh, M. Kulkarni, and S. Roy, Eigenvector correlations across the localization transition in non-Hermitian power-law banded random matrices, Phys. Rev. B 108, L060201 (2023).
- Gong et al. [2018] Z. Gong, Y. Ashida, K. Kawabata, K. Takasan, S. Higashikawa, and M. Ueda, Topological Phases of Non-Hermitian Systems, Phys. Rev. X 8, 031079 (2018).
- Zhou and Lee [2019] H. Zhou and J. Y. Lee, Periodic table for topological bands with non-Hermitian symmetries, Phys. Rev. B 99, 235112 (2019).
- Porras and Fernández-Lorenzo [2019] D. Porras and S. Fernández-Lorenzo, Topological Amplification in Photonic Lattices, Phys. Rev. Lett. 122, 143901 (2019).
- Ramos et al. [2022] T. Ramos, Álvaro Gómez-León, J. J. García-Ripoll, A. González-Tudela, and D. Porras, Directional Josephson traveling-wave parametric amplifier via non-Hermitian topology, arXiv:2207.13728 (2022).
- Herviou et al. [2019] L. Herviou, J. H. Bardarson, and N. Regnault, Defining a bulk-edge correspondence for non-Hermitian Hamiltonians via singular-value decomposition, Phys. Rev. A 99, 052118 (2019).
- Wanjura et al. [2020] C. C. Wanjura, M. Brunelli, and A. Nunnenkamp, Topological framework for directional amplification in driven-dissipative cavity arrays, Nat. Commun. 11, 3149 (2020).
- Brunelli et al. [2022] M. Brunelli, C. C. Wanjura, and A. Nunnenkamp, Restoration of the non-Hermitian bulk-boundary correspondence via topological amplification, arXiv:2207.12427 (2022).
- Datta [1995] S. Datta, Electronic Transport in Mesoscopic Systems (Cambridge University Press, Cambridge, England, 1995).
- Imry [1997] Y. Imry, Introduction to Mesoscopic Physics (Oxford University Press, New York, 1997).
- Lee and Stone [1985] P. A. Lee and A. D. Stone, Universal Conductance Fluctuations in Metals, Phys. Rev. Lett. 55, 1622 (1985).
- Lee et al. [1987] P. A. Lee, A. D. Stone, and H. Fukuyama, Universal conductance fluctuations in metals: Effects of finite temperature, interactions, and magnetic field, Phys. Rev. B 35, 1039 (1987).
- Al’tshuler [1985] B. L. Al’tshuler, Fluctuations in the extrinsic conductivity of disordered conductors, JETP Lett. 41, 648 (1985).
- Imry [1986] Y. Imry, Active Transmission Channels and Universal Conductance Fluctuations, Europhys. Lett. 1, 249 (1986).
- Al’tshuler and Shklovskiĭ [1986] B. L. Al’tshuler and B. I. Shklovskiĭ, Repulsion of energy levels and conductivity of small metal samples, J. Exp. Theor. Phys 64, 127 (1986).
- Mostafazadeh [2002a] A. Mostafazadeh, Pseudo-Hermiticity versus PT symmetry: The necessary condition for the reality of the spectrum of a non-Hermitian Hamiltonian, J. Math. Phys. 43, 205 (2002a).
- Mostafazadeh [2002b] A. Mostafazadeh, Pseudo-Hermiticity versus PT symmetry II: A complete characterization of non-Hermitian Hamiltonians with a real spectrum, J. Math. Phys. 43, 2814 (2002b).
- Oganesyan and Huse [2007] V. Oganesyan and D. A. Huse, Localization of interacting fermions at high temperature, Phys. Rev. B 75, 155111 (2007).
- Atas et al. [2013] Y. Y. Atas, E. Bogomolny, O. Giraud, and G. Roux, Distribution of the Ratio of Consecutive Level Spacings in Random Matrix Ensembles, Phys. Rev. Lett. 110, 084101 (2013).
- Ginibre [1965] J. Ginibre, Statistical Ensembles of Complex, Quaternion, and Real Matrices, J. Math. Phys. 6, 440 (1965).
- Girko [1985] V. L. Girko, Circular Law, Theory Probab. Appl. 29, 694 (1985).
- Mehta [2004] M. L. Mehta, Random Matrices (Elsevier, Amsterdam, 2004).
- Byun and Forrester [2022] S.-S. Byun and P. J. Forrester, Progress on the study of the Ginibre ensembles I: GinUE, arXiv:2211.16223 (2022).
- Byun and Forrester [2023] S.-S. Byun and P. J. Forrester, Progress on the study of the Ginibre ensembles II: GinOE and GinSE, arXiv:2301.05022 (2023).
- Edelman [1988] A. Edelman, Eigenvalues and Condition Numbers of Random Matrices, SIAM J. Matrix Anal. Appl. 9, 543 (1988).
- Forrester [1993] P. J. Forrester, The spectrum edge of random matrix ensembles, Nucl. Phys. B 402, 709 (1993).
- Nishigaki et al. [1998] S. M. Nishigaki, P. H. Damgaard, and T. Wettig, Smallest Dirac eigenvalue distribution from random matrix theory, Phys. Rev. D 58, 087704 (1998).
- Damgaard and Nishigaki [2001] P. H. Damgaard and S. M. Nishigaki, Distribution of the smallest Dirac operator eigenvalue, Phys. Rev. D 63, 045012 (2001).
- Sun and Ye [2020] F. Sun and J. Ye, Periodic Table of the Ordinary and Supersymmetric Sachdev-Ye-Kitaev Models, Phys. Rev. Lett. 124, 244101 (2020).
- Kawabata et al. [2019b] K. Kawabata, S. Higashikawa, Z. Gong, Y. Ashida, and M. Ueda, Topological unification of time-reversal and particle-hole symmetries in non-Hermitian physics, Nat. Commun. 10, 297 (2019b).
- Gorini et al. [1976] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Completely positive dynamical semigroups of ‐level systems, J. Math. Phys. 17, 821 (1976).
- Lindblad [1976] G. Lindblad, On the generators of quantum dynamical semigroups, Commun. Math. Phys. 48, 119 (1976).
- Hamazaki and Ueda [2019] R. Hamazaki and M. Ueda, Random-matrix behavior of quantum nonintegrable many-body systems with Dyson’s three symmetries, Phys. Rev. E 99, 042116 (2019).
- V. A. Marčenko and L. A. Pastur [1967] V. A. Marčenko and L. A. Pastur, Distribution of eigenvalues for some sets of random matrices, Math. USSR-Sbornik 1, 457 (1967).
- Oas [1997] G. Oas, Universal cubic eigenvalue repulsion for random normal matrices, Phys. Rev. E 55, 205 (1997).
- Chau and Zaboronsky [1998] L.-L. Chau and O. Zaboronsky, On the Structure of Correlation Functions in the Normal Matrix Model, Commun. Math. Phys. 196, 203 (1998).
- Elbau and Felder [2005] P. Elbau and G. Felder, Density of Eigenvalues of Random Normal Matrices, Commun. Math. Phys. 259, 433 (2005).
- Mahaux and Weidenmüller [1968] C. Mahaux and H. A. Weidenmüller, Comparison between the -Matrix and Eigenchannel Methods, Phys. Rev. 170, 847 (1968).
- Mahaux and Weidenmüller [1969] C. Mahaux and H. A. Weidenmüller, Shell-Model Approach to Nuclear Reactions (North-Holland Publishing Company, Amsterdam, 1969).
- Verbaarschot et al. [1985] J. J. M. Verbaarschot, H. A. Weidenmüller, and M. R. Zirnbauer, Grassmann integration in stochastic quantum physics: The case of compound-nucleus scattering, Phys. Rep. 129, 367 (1985).
- Zirnstein et al. [2021] H.-G. Zirnstein, G. Refael, and B. Rosenow, Bulk-Boundary Correspondence for Non-Hermitian Hamiltonians via Green Functions, Phys. Rev. Lett. 126, 216407 (2021).
- Zirnstein and Rosenow [2021] H.-G. Zirnstein and B. Rosenow, Exponentially growing bulk Green functions as signature of nontrivial non-Hermitian winding number in one dimension, Phys. Rev. B 103, 195157 (2021).