This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\UseRawInputEncoding

Complex Berry phase and imperfect non-Hermitian phase transitions

Stefano Longhi Dipartimento di Fisica, Politecnico di Milano, Piazza L. da Vinci 32, I-20133 Milano, Italy IFISC (UIB-CSIC), Instituto de Fisica Interdisciplinar y Sistemas Complejos, E-07122 Palma de Mallorca, Spain [email protected]    Liang Feng Department of Materials Science and Engineering, University of Pennsylvania, Philadelphia, PA 19104, USA
Abstract

In many classical and quantum systems described by an effective non-Hermitian Hamiltonian, spectral phase transitions, from an entirely real energy spectrum to a complex spectrum, can be observed as a non-Hermitian parameter in the system is increased above a critical value. A paradigmatic example is provided by systems possessing parity-time (𝒫𝒯\mathcal{PT}) symmetry, where the energy spectrum remains entirely real in the unbroken 𝒫𝒯\mathcal{PT} phase while a transition to complex energies is observed in the unbroken 𝒫𝒯\mathcal{PT} phase. Such spectral phase transitions are universally sharp. However, when the system is slowly and periodically cycled, the phase transition can become smooth, i.e. imperfect, owing to the complex Berry phase associated to the cyclic adiabatic evolution of the system. This remarkable phenomenon is illustrated by considering the spectral phase transition of the Wannier-Stark ladders in a 𝒫𝒯\mathcal{PT}-symmetric class of two-band non-Hermitian lattices subjected to an external dc field, unraveling that a non-vanishing imaginary part of the Zak phase - the Berry phase picked up by a Bloch eigenstate evolving across the entire Brillouin zone- is responsible for imperfect spectral phase transitions.

I Introduction

The geometric or Berry phase b1 ; b2 ; b3 ; b3bis , a concept which was systematized and popularized in the 1980s by Sir Michael Berry b1 , has permeated through all branches of physics with applications in diverse fields ranging from atomic and molecular physics b4 ; b5 ; b6 to condensed-matter physics b7 ; b8 ; b9 , classical optics b10 ; b11 ; b12 , high energy and particle physics b13 ; b14 ; b15 , gravity and cosmology b16 . When a quantum or classical system undergoes a cyclic evolution governed by a change of parameters, besides the dynamical phase it acquires an additional phase term, the Berry phase, which depends only on the geometry of the path but not on how the cycle is run. In condensed matter physics, the geometric phase manifests itself in many phenomena, such as the quantum Hall effect, electric polarization, orbital magnetism and exchange statistics b3 ; b8 ; b9 . In a crystal, the application of an electric field changes the quasi-momentum of the electronic wave function over the entire Brillouin zone, and the accumulated geometric phase is known as the Zak phase b17 . In one-dimensional (1D) lattices, the bulk topological properties of the Bloch bands are characterized by the quantized Zak phase b18 ; b19 ; b20 ; b21 ; b22 ; b23 ; b23bis , which can serve as a topological number.

Several exciting phenomena that are attracting great interest in modern condensed-matter physics and beyond, such as non-Hermitian skin effect, modified bulk-boundary correspondence, exceptional points, nontrivial spectral topology and phase transitions, etc. b24 ; b25 ; b26 ; b27 ; b27bis , appear in non-Hermitian models, i.e. in models where the dynamics is described by an effective non-Hermitian Hamiltonian b28 ; b29 ; b30 which accounts for energy/particle exchange with external reservoirs. A remarkable property of certain classes of non-Hermitian Hamiltonians is to display an entirely real energy spectrum in spite of non-Hermiticity b31 ; b32 ; b33 ; b34 ; b35 ; b36 ; b37 ; b38 ; b39 ; b39bis ; b40 . Among such Hamiltonians, great attention has been devoted to the ones displaying parity-time (𝒫𝒯\mathcal{PT}) symmetry b31 ; b32 ; b33 , a concept that has become very popular in the past decade and found important applications in photonics and beyond b41 ; b42 ; b43 ; b44 ; b45 ; b46 ; b47 . For given parity 𝒫\mathcal{P} and time reversal 𝒯\mathcal{T} operators, an Hamiltonian \mathcal{H} is said to be 𝒫𝒯\mathcal{PT}-symmetric if the commutator [,𝒫𝒯][\mathcal{H},\mathcal{PT]} vanishes, i.e. 𝒫𝒯=𝒫𝒯\mathcal{HPT}=\mathcal{PTH}. However, since the operator 𝒫𝒯\mathcal{PT} is not linear, 𝒫𝒯\mathcal{PT}-symmetry itself does not necessarily imply that the \mathcal{H} and 𝒫𝒯\mathcal{PT} operators share the same set of eigenfunctions. This means that, while the underlying Hamiltonian \mathcal{H} possesses 𝒫𝒯\mathcal{PT} symmetry, i.e. 𝒫𝒯=𝒫𝒯\mathcal{PTH}=\mathcal{HPT}, the corresponding eigenfunctions |E|E\rangle of \mathcal{H} can (or cannot) display the same symmetry. When some eigenfunctions of \mathcal{H} break the 𝒫𝒯\mathcal{PT} symmetry, i.e. 𝒫𝒯|E\mathcal{PT}|E\rangle and |E|E\rangle are distinct states, we have a typical scenario of spontaneous symmetry breaking b33 . Spontaneous 𝒫𝒯\mathcal{PT} symmetry breaking corresponds to a spectral phase transition, from an entirely real energy spectrum in the unbroken 𝒫𝒯\mathcal{PT} phase to a complex energy spectrum in the spontaneously broken 𝒫𝒯\mathcal{PT} phase. When a control parameter in the system is varied above a critical value, spontaneous 𝒫𝒯\mathcal{PT} symmetry breaking is usually observed and in the broken 𝒫𝒯\mathcal{PT} phase energies appear in complex conjugate pairs. This readily follows from the anti-linear nature of the 𝒯\mathcal{T} operator: if |E|E\rangle is an eigenfunction of \mathcal{H} with eigenenergy EE, i.e. |E=E|E\mathcal{H}|E\rangle=E|E\rangle, then 𝒫𝒯|E=𝒫𝒯|E=𝒫𝒯E|E=E𝒫𝒯|E\mathcal{HPT}|E\rangle=\mathcal{PTH}|E\rangle=\mathcal{PT}E|E\rangle=E^{*}\mathcal{PT}|E\rangle. This means that 𝒫𝒯|E\mathcal{PT}|E\rangle is an eigenfunction of \mathcal{H} with eigenenergy EE^{*}. When the symmetry is not spontaneously broken, |E|E\rangle and 𝒫𝒯|E\mathcal{PT}|E\rangle are the same eigenfunction, which necessarily implies E=EE=E^{*}: in the unbroken 𝒫𝒯\mathcal{PT} phase the energy spectrum is entirely real. On the other hand, when the symmetry is spontaneously broken, |E|E\rangle is not necessarily an eigenfunction of the 𝒫𝒯\mathcal{PT} operator, and thus the eigenfunctions |E|E\rangle and 𝒫𝒯|E\mathcal{PT}|E\rangle, with non-degenerate eigenenergies EE and EE^{*}, are linearly independent: in this case the energy spectrum becomes complex and formed by complex conjugate pairs. The spontaneous symmetry breaking phase transition is ubiquitously sharp and the symmetry breaking point corresponds to the appearance of non-Hermitian degeneracies, i.e. exceptional points b47 ; b47bis ; b59 or spectral singularities sp1 ; sp2 ; sp3 , at the critical point.
The concept of geometric phase can be generalized to non-Hermitian systems, providing a geometrical description of the quantum evolution of non-Hermitian systems under cyclic variation of parameters b48 ; b49 ; b50 ; b51 ; b52 ; b53 ; b54 ; new1 ; b54noo ; b54basta1 ; b54basta2 ; b54bis ; b55 ; b56 ; b55bis ; b57 . As compared to Hermitian systems, different forms of Berry phases have been introduced. Here we will use the Berry phase from the biorthogonal basis of the non-Hermitian Hamiltonain, which is thus rather generally complex. An interesting property of adiabatic cycling in non-Hermitian systems is that the energy surface can display a nontrivial topology: when one follows a loop in the space of system parameters, even in the absence of degeneracies the energies and corresponding instantaneous eigenstates may swap places, which renders the evolution non-cyclic. The interchange of energies arises when exceptional points are encircled in the space of system parameters b47 ; b58 ; b59 . The complex Berry phase has been suggested to provide a topological invariant identifying different topological phases and quantum phase transitions in certain non-Hermitian models b27bis ; b60 ; b61 ; b62 ; b62bis ; b63 ; b63bis ; b64 ; b65 ; b66 ; b66bis ; b66tris ; b67 ; b68 ; b69 ; b70 , and some general conditions for the quantization of the Berry phase under certain generalized symmetries have been provided b71 . In a non-Hermitian lattice, complex Berry phase, i.e. Zak phase, naturally arises under an external dc force or a time-varying magnetic flux b65 , so that a Bloch eigenstate adiabatically evolves across the entire Brillouin zone accumulating a complex geometric phase. While the related phenomena of Bloch oscillations and Zener tunneling have been investigated at some extent in non-Hermitian lattices b72 ; b72bis ; b73 ; b74 ; b75 ; b76 ; b77 ; b78 ; b79 ; b80 ; b81 , physical signatures of the complex Zak phase have received so far little attention and mostly restricted to some specific lattice models b54basta1 ; b65 .

In this work we show that the complex Berry phase in slow-cycled non-Hermitian 𝒫𝒯\mathcal{PT} symmetric systems can lead to imperfect, i.e. smooth, spectral phase transitions. This phenomenon is first illustrated by considering a general model of two-level 𝒫𝒯\mathcal{PT} symmetric systems, and then applied to explain the imperfect phase transition of Wannier-Stark ladders found in certain two-band non-Hermitian lattices b81 , which is rooted in the non-vanishing imaginary part of the Zak phase.

II Phase transitions in a cycled two-level 𝒫𝒯\mathcal{PT} symmetric model

II.1 Model and 𝒫𝒯\mathcal{PT} symmetry breaking phase transition

We consider a classical or quantum two-level system described by an effective 2×22\times 2 non-Hermitian matrix Hamiltonian =(k)\mathcal{H}=\mathcal{H}(k), which depends on a real parameter kk and is periodic in kk with a period of 2π2\pi, i.e. (k+2π)=(k)\mathcal{H}(k+2\pi)=\mathcal{H}(k). As we will discuss in the next section, in the Wannier-Stark ladder problem of a non-Hermitian lattice driven by a dc field the matrix Hamiltonian (k)\mathcal{H}(k) corresponds to the Bloch Hamiltonian of a two-band lattice and kk is the quasi momentum, that drifts to span the entire Brillouin zone in the presence of the dc field.
The temporal dynamics of the system is described by the Schrödinger equation

iddt(ψ1ψ2)=(11122122)(ψ1ψ2)=(k)(ψ1ψ2).i\frac{d}{dt}\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\left(\begin{array}[]{cc}\mathcal{H}_{11}&\mathcal{H}_{12}\\ \mathcal{H}_{21}&\mathcal{H}_{22}\end{array}\right)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\mathcal{H}(k)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right). (1)

We assume that the Hamiltonian is 𝒫𝒯\mathcal{PT} symmetric with parity 𝒫\mathcal{P} and time reversal 𝒯\mathcal{T} operators defined by

𝒫=σx=(0110),𝒯=𝒦\mathcal{P}=\sigma_{x}=\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right)\;,\;\;\mathcal{T}=\mathcal{K} (2)

where σx\sigma_{x} is the Pauli matrix and 𝒦\mathcal{K} the element-wise complex conjugation operator. 𝒫𝒯\mathcal{PT} symmetry, i.e. the condition 𝒫𝒯=𝒫𝒯\mathcal{PTH}=\mathcal{HPT}, is satisfied provided that

22=11,21=12\mathcal{H}_{22}=\mathcal{H}^{*}_{11}\;,\;\;\mathcal{H}_{21}=\mathcal{H}^{*}_{12} (3)

so that the non-Hermiticity in the system is embedded in a non-vanishing imaginary part of 11\mathcal{H}_{11}. The most general form of matrix elements that respect the 𝒫𝒯\mathcal{PT} symmetry is thus

11=22=G(k)+iλW(k)\displaystyle\mathcal{H}_{11}=\mathcal{H}_{22}^{*}=G(k)+i\lambda W(k) (4)
12=21=R(k)exp[iφ(k)]\displaystyle\mathcal{H}_{12}=\mathcal{H}_{21}^{*}=R(k)\exp[i\varphi(k)] (5)

where G(k),W(k)G(k),W(k), R(k)R(k) are real and periodic functions of kk with period 2π2\pi, φ(k)\varphi(k) is a real function with φ(k+2π)=φ(k)\varphi(k+2\pi)=\varphi(k) mod 2π2\pi, and λ0\lambda\geq 0 is a real parameter that measures the strength of non-Hermiticity in the system, the case λ=0\lambda=0 corresponding to (k)\mathcal{H}(k) Hermitian. Further, we assume that R(k)R(k) is nonvanishing over the entire range 0k2π0\leq k\leq 2\pi.
When the parameter kk is kept constant, the eigenenergies of (k)\mathcal{H}(k) are given by

E±(k)=G(k)±R2(k)λ2W2(k)E_{\pm}(k)=G(k)\pm\sqrt{R^{2}(k)-\lambda^{2}W^{2}(k)} (6)

with corresponding (right) eigenvectors

𝐮+(k)\displaystyle\mathbf{u}_{+}(k) =\displaystyle= (cos(θ2)sin(θ2)exp(iφ))\displaystyle\left(\begin{array}[]{c}\cos\left(\frac{\theta}{2}\right)\\ \sin\left(\frac{\theta}{2}\right)\exp(-i\varphi)\end{array}\right) (9)
𝐮(k)\displaystyle\mathbf{u}_{-}(k) =\displaystyle= (sin(θ2)cos(θ2)exp(iφ)).\displaystyle\left(\begin{array}[]{c}\sin\left(\frac{\theta}{2}\right)\\ -\cos\left(\frac{\theta}{2}\right)\exp(-i\varphi)\end{array}\right). (12)

In the previous equations, the complex angle θ=θ(k)\theta=\theta(k) is defined by the relation

tanθ(k)=R(k)iλW(k).\tan\theta(k)=\frac{R(k)}{i\lambda W(k)}. (13)

Note that the imaginary part of the angle θ(k)\theta(k) diverges when R(k)=±λW(k)R(k)=\pm\lambda W(k), corresponding to the simultaneous coalescence of the two energies and eigenstates, i.e. to the appearance of an exceptional point. The left eigenvectors of (k)\mathcal{H}(k), i.e. the (right) eigenvectors of the adjoint (k)\mathcal{H}^{{\dagger}}(k) with eigenvalues E±(k)E_{\pm}^{*}(k), read

𝐯+(k)\displaystyle\mathbf{v}_{+}(k) =\displaystyle= (cos(θ2)sin(θ2)exp(iφ))\displaystyle\left(\begin{array}[]{c}\cos^{*}\left(\frac{\theta}{2}\right)\\ \sin^{*}\left(\frac{\theta}{2}\right)\exp(-i\varphi)\end{array}\right) (16)
𝐯(k)\displaystyle\mathbf{v}_{-}(k) =\displaystyle= (sin(θ2)cos(θ2)exp(iφ))\displaystyle\left(\begin{array}[]{c}\sin^{*}\left(\frac{\theta}{2}\right)\\ -\cos^{*}\left(\frac{\theta}{2}\right)\exp(-i\varphi)\end{array}\right) (19)

and the biorthogonal conditions

𝐯n(k)|𝐮m(k)=δn,m\langle\mathbf{v}_{n}(k)|\mathbf{u}_{m}(k)\rangle=\delta_{n,m} (20)

are satisfied for any kk, with n,m=+,n,m=+,-. After letting

λc(k)|R(k)/W(k)|,\lambda_{c}(k)\equiv|R(k)/W(k)|, (21)

from Eq.(6) it readily follows that the energy spectrum is real for λ<λc(k)\lambda<\lambda_{c}(k) (unbroken 𝒫𝒯\mathcal{PT} phase), and complex for λ>λc(k)\lambda>\lambda_{c}(k) (broken 𝒫𝒯\mathcal{PT} phase), with the appearance of an exceptional point at the critical point λ=λc(k)\lambda=\lambda_{c}(k).

II.2 Phase transition in the cycled system

Let us now consider the two-level system when the Hamiltonian (k)\mathcal{H}(k) is periodically and adiabatically cycled in time. We assume that the parameter kk in the Hamiltonian varies in time according to

k=ωtk=\omega t (22)

where ω\omega is the cycling frequency. The temporal dynamics of the two-level system is thus described by the equation

iddt(ψ1ψ2)=(ωt)(ψ1ψ2).i\frac{d}{dt}\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\mathcal{H}(\omega t)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right). (23)

In most of our analysis, we will limit our attention considering the system dynamics in the slow-cycling regime ω0\omega\rightarrow 0. As we will comment below with reference to some specific examples, the main motivation thereof is that to observe smooth spectral phase transitions the system evolution must be slow enough. Without loss of generality we can assume G(k)=0G(k)=0, i.e. 11(k)=22(k)=iλW(k)\mathcal{H}_{11}(k)=\mathcal{H}_{22}^{*}(k)=i\lambda W(k), so that the instantaneous eigenenergies of (k)\mathcal{H}(k) read

E±(k)=±R2(k)λ2W2(k)E_{\pm}(k)=\pm\sqrt{R^{2}(k)-\lambda^{2}W^{2}(k)} (24)

with k=ωtk=\omega t. In fact, a non-vanishing value of G(k)G(k) can be eliminated from the dynamics after the gauge transformation

(ψ1(t)ψ2(t))(ψ1(t)ψ2(t))exp{iω0ωtG(k)𝑑k}.\left(\begin{array}[]{c}\psi_{1}(t)\\ \psi_{2}(t)\end{array}\right)\rightarrow\left(\begin{array}[]{c}\psi_{1}(t)\\ \psi_{2}(t)\end{array}\right)\exp\left\{-\frac{i}{\omega}\int_{0}^{\omega t}G(k)dk\right\}.

According to Floquet theory, the most general solution to the Schrödinger equation (15) is given by

(ψ1(t)ψ2(t))=𝒰(t)exp(it)(ψ1(0)ψ2(0))\left(\begin{array}[]{c}\psi_{1}(t)\\ \psi_{2}(t)\end{array}\right)=\mathcal{U}(t)\exp(-i\mathcal{R}t)\left(\begin{array}[]{c}\psi_{1}(0)\\ \psi_{2}(0)\end{array}\right) (25)

where \mathcal{R} is a time-independent 2×22\times 2 matrix while 𝒰(t)\mathcal{U}(t) is a time-dependent and periodic 2×22\times 2 matrix, 𝒰(t+2π/ω)=𝒰(t)\mathcal{U}(t+2\pi/\omega)=\mathcal{U}(t), with 𝒰(0)=\mathcal{U}(0)=\mathcal{I} (the identity matrix). The exponential of the matrix \mathcal{R} can be expressed in terms of the path-ordered integral

exp(i)=𝒯¯exp[iω2π02π/ω𝑑t(ωt)]\exp(-i\mathcal{R})=\bar{\mathcal{T}}\exp\left[-i\frac{\omega}{2\pi}\int_{0}^{2\pi/\omega}dt\;\mathcal{H}(\omega t)\right]

where 𝒯¯\bar{\mathcal{T}} indicates the time ordering. The two eigenvalues μ±=μ±(λ)\mu_{\pm}=\mu_{\pm}(\lambda) of \mathcal{R} are the quasi energies of the time-periodic cycled system. The real parts of the quasi energies are defined apart from integer multiples than ω\omega. Note that for G(k)=0G(k)=0 the trace of (k)\mathcal{H}(k) vanishes, so that μ=μ+\mu_{-}=-\mu_{+}, i.e. the two quasi energies can be assumed to be opposite one another. A non-vanishing value of G(k)G(k) would just lead to a shift of the quasi energies by the amount (1/2π)02π𝑑kG(k)(1/2\pi)\int_{0}^{2\pi}dkG(k).
A natural question arises: akin to the non-cycled 𝒫𝒯\mathcal{PT} symmetric system, is there a spectral phase transition, from real to complex quasi energies, as the non-Hermitian parameter λ\lambda in the system is increased above a critical value? To answer this question, let us indicate by λ¯c\bar{\lambda}_{c} the minimum value of λc(k)\lambda_{c}(k) as kk spans the range 0k2π0\leq k\leq 2\pi, i.e.

λ¯c=min0k2πλc(k)=min0k2π|R(k)W(k)|.\bar{\lambda}_{c}=\min_{0\leq k\leq 2\pi}\lambda_{c}(k)=\min_{0\leq k\leq 2\pi}\left|\frac{R(k)}{W(k)}\right|. (26)

Intuitively, for a slowly-cycled system one would expect the following scenario: for λ<λ¯c\lambda<\bar{\lambda}_{c}, the instantaneous eigenenergies E±(k=ωt)E_{\pm}(k=\omega t) of (k=ωt)\mathcal{H}(k=\omega t) are real, and thus we expect the quasi energies μ±\mu_{\pm} to remain real as well. On the other hand, for λ>λ¯c\lambda>\bar{\lambda}_{c} within the modulation cycle there are time intervals where the instantaneous eigenenergies E±(k=ωt)E_{\pm}(k=\omega t) become complex: in this case we expect the quasi energies to become complex too. Hence, according to such an intuitive picture, we expect a spectral phase transition of the cycled two-level system, from real to complex quasi energies, when the non-Hermitian parameter λ\lambda is increased above the critical value λ¯c\bar{\lambda}_{c}. This result is indeed what one observes from a numerical computation of the quasi energies in several examples of cycled two-level 𝒫𝒯\mathcal{PT} symmetric models, as shown in the next subsection. However, in some other models it turns out that, for a small but non-vanishing oscillation frequency ω\omega, the phase transition is smooth, i.e. imperfect: below the critical value λ¯c\bar{\lambda}_{c} the imaginary part of the quasi energy takes a small but non-vanishing value, which scales as ω\sim\omega, i.e. it exactly vanishes only in the limit ω0\omega\rightarrow 0. What is the physical origin of such an imperfect phase transition, which is observed in some models but not in others?
The answer to this question is rooted in the appearance of a complex Berry phase in certain models (but not in others), and can be gained from an adiabatic analysis of the time evolution of the system in the ω0\omega\rightarrow 0 limit, which is detailed in the Appendices A and B. In the adiabatic analysis, the slow evolution of the amplitudes of the instantaneous eigenstates 𝐮±(k=ωt)\mathbf{u}_{\pm}(k=\omega t) of the Hamiltonian (k=ωt)\mathcal{H}(k=\omega t) is governed by the non-Hermitian Berry connection

𝒜n,l(k)=i𝐯n|k𝐮l\mathcal{A}_{n,l}(k)=-i\langle\mathbf{v}_{n}|\partial_{k}\mathbf{u}_{l}\rangle (27)

(n,l=+,n,l=+,-), which is defined in the context of the biorthonormal inner product. The integrals of the diagonal terms of Berry connection, 𝒜+,+(k)\mathcal{A}_{+,+}(k) and 𝒜,(k)\mathcal{A}_{-,-}(k), over the interval 0k2π0\leq k\leq 2\pi, i.e.

γB+=02π𝑑k𝒜+,+(k),γB=02π𝑑k𝒜,(k)\gamma_{B_{+}}=\int_{0}^{2\pi}dk\mathcal{A}_{+,+}(k)\;,\;\;\gamma_{B_{-}}=\int_{0}^{2\pi}dk\mathcal{A}_{-,-}(k) (28)

are the non-Hermitian Berry phases associated to the two instantaneous eigenstates 𝐮±(k)\mathbf{u}_{\pm}(k). The explicit form of the Berry connection and Berry phases are derived in Appendix A. In particular, one has

γB±=1202π𝑑kdφdk±i202π𝑑kdφdksinhψ(k)\gamma_{B_{\pm}}=\mp\frac{1}{2}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\pm\frac{i}{2}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\sinh\psi(k) (29)

where the function ψ(k)\psi(k) is defined by the relation

tanhψ(k)=λW(k)R(k).\tanh\psi(k)=\frac{\lambda W(k)}{R(k)}. (30)

Note that the Berry phase vanishes in any 𝒫𝒯\mathcal{PT}-symmetric two-level system with (dφ/dk)0(d\varphi/dk)\equiv 0.
The adiabatic analysis shows some subtleties and limitations when applied to our model, owing to the appearance of instantaneous EP on the cycle when λ>λc¯\lambda>\bar{\lambda_{c}}. Technical details are given in Appendix B. The main result of the adiabatic analysis is that, in the limit ω0\omega\rightarrow 0 and for λλc¯\lambda\neq\bar{\lambda_{c}}, the two quasi energies are given by

μ±=12π02π𝑑kE±(k)+ω2πγB±.\mu_{\pm}=\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{\pm}(k)+\frac{\omega}{2\pi}\gamma_{B_{\pm}}. (31)

Note that, since E(k)=E+(k)E_{-}(k)=-E_{+}(k) and γB=γB+\gamma_{B_{-}}=-\gamma_{B_{+}}, one has μ=μ+\mu_{-}=-\mu_{+}, as it should. The above result provides an approximate form of the quasi energies in the adiabatic limit ω0\omega\rightarrow 0 for any strength λ\lambda of the non-Hermitian parameter far from the critical value λ=λc¯\lambda=\bar{\lambda_{c}}, at which the Berry phase term becomes singular and the adiabatic analysis fails; a discussion on this point is given in the Appendix B.
According to Eq.(23), each quasi energy is given by the sum of two terms. The first one is related to the dynamical phase accumulated by the adiabatic eigenstates in one cycle and equals the average of the instantaneous energies E±(k)E_{\pm}(k) over one cycle. The dynamical phase term is clearly independent of the modulation frequency ω\omega and is real for λ<λ¯c\lambda<\bar{\lambda}_{c}, while its imaginary part is non-vanishing for λ>λ¯c\lambda>\bar{\lambda}_{c}. The second term on the right hand side in Eq.(23) is the non-Hermitian Berry phase contribution. This is a small term which vanishes like ω\sim\omega as ω0\omega\rightarrow 0. Interestingly, for λ<λ¯c\lambda<\bar{\lambda}_{c} the imaginary part of the quasi energies is provided solely by the Berry phase term, and vanishes as ω0\omega\rightarrow 0. This explains why in the cycled two-level 𝒫𝒯\mathcal{PT}-symmetric system with a vanishing Berry phase the spectral phase transition of the quasi energies, at λ=λ¯c\lambda=\bar{\lambda}_{c}, is sharp (exact), while it becomes smooth (imperfect) when the non-Hermitian Berry phase along the cycle is non-vanishing.

II.3 Illustrative examples

The main result of the adiabatic analysis is that the spectral phase transition of the quasi energies in the slow-cycled 𝒫𝒯\mathcal{PT}-symmetric two-level system turns out be be imperfect (smooth) whenever the Berry phase in the cycle is complex, which requires the derivative (dφ/dk)(d\varphi/dk) not to identically vanish. On the other hand, the phase transition is sharp (exact) whenever the Berry phase is real. Here we illustrate and confirm the predictions of the adiabatic analysis by considering three examples of cycled two-level 𝒫𝒯\mathcal{PT}-symmetric systems.

1. First example. The first example is a simple and exactly-solvable model, corresponding to G(k)=0G(k)=0, W(k)=1W(k)=1, R(k)=R0R(k)=R_{0}, φ(k)=k\varphi(k)=k, i.e. to the 𝒫𝒯\mathcal{PT}-symmetric Hamiltonian

(k)=(iλR0exp(ik)R0exp(ik)iλ)\mathcal{H}(k)=\left(\begin{array}[]{cc}i\lambda&R_{0}\exp(ik)\\ R_{0}\exp(-ik)&-i\lambda\end{array}\right) (32)

where R0>0R_{0}>0 is a real parameter. Physically, this model describes a two-level system, in which the two states are Hermitian-coupled by an amplitude R0R_{0} and a gauge (Peierls) phase kk and with gain (λ\lambda) and loss (λ-\lambda) rates in the two levels. The system displays a 𝒫𝒯\mathcal{PT} symmetry breaking at a critical value λc(k)=λ¯c\lambda_{c}(k)=\bar{\lambda}_{c}, independent of kk, given by

λ¯c=R0.\bar{\lambda}_{c}=R_{0}. (33)

In the cycled system with k=ωtk=\omega t, we expect an imperfect phase transition of quasi energies because (dφ/dk)=10(d\varphi/dk)=1\neq 0 and the imaginary part of the Berry phase does not vanish. The quasi energies μ±\mu_{\pm} can be calculated in an exact form, given that the time dependence of the Hamiltonian (k=ωt)\mathcal{H}(k=\omega t) can be removed from the dynamics after the gauge transformation

ψ1(t)=ψ¯1(t)exp(iωt2),ψ2(t)=ψ¯2(t)exp(iωt2).\psi_{1}(t)=\bar{\psi}_{1}(t)\exp\left(i\frac{\omega t}{2}\right)\;,\;\psi_{2}(t)=\bar{\psi}_{2}(t)\exp\left(-i\frac{\omega t}{2}\right). (34)

The exact expression of the quasi energies can be readily computed, yielding

μ±=±R02+(ω2+iλ)2ω2\mu_{\pm}=\pm\sqrt{R_{0}^{2}+\left(\frac{\omega}{2}+i\lambda\right)^{2}}\mp\frac{\omega}{2} (35)

A typical behavior of the imaginary parts of the quasi energies versus λ\lambda, in the adiabatic limit ωR0\omega\ll R_{0}, is depicted in Fig.1, clearly showing the appearance of an imperfect spectral phase transition near λ=λ¯c\lambda=\bar{\lambda}_{c}. Note that, when λ\lambda is not too close to λ¯c=R0\bar{\lambda}_{c}=R_{0}, in the adiabatic limit ω0\omega\rightarrow 0 we can expand the right hand side of Eq.(27) in power series of ω\omega and, up to first order in ω\omega, the following approximate expression of quasi energies is obtained

μ±±R02λ2ω2±iλω2R02λ2.\mu_{\pm}\simeq\pm\sqrt{R_{0}^{2}-\lambda^{2}}\mp\frac{\omega}{2}\pm i\frac{\lambda\omega}{2\sqrt{R_{0}^{2}-\lambda^{2}}}. (36)

It can be readily shown that Eq.(28) precisely reproduces the result predicted by the adiabatic analysis [Eq.(23)]. In fact, the dynamical phase contribution to the quasi energy is given by

12π02π𝑑kE±(k)=±R02λ2\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{\pm}(k)=\pm\sqrt{R_{0}^{2}-\lambda^{2}}
Refer to caption
Figure 1: Behavior of the imaginary part of the quasi energies μ±\mu_{\pm} versus the non-Hermitian parameter λ\lambda for the cycled 𝒫𝒯\mathcal{PT}-symmetric system with Hamiltonian (k)\mathcal{H}(k) given by Eq.(24) for R0=1R_{0}=1 and for a modulation frequency (a) ω=0.02\omega=0.02, and (b) ω=0.1\omega=0.1. Open blue circles and red crosses refer to the exact curves and to the approximate curves obtained from the adiabatic analysis, respectivley. Note that the spectral phase transition is imperfect around the critical point λ=λ¯c=R0\lambda=\bar{\lambda}_{c}=R_{0}, and that near the critical point the adiabatic curves fail to predict the exact behavior of the quasi energies.

while the Berry phase contribution reads

ω2πγB±\displaystyle\frac{\omega}{2\pi}\gamma_{B_{\pm}} =\displaystyle= ω4π02π𝑑kdφdk±iω4π02π𝑑kdφdksinhψ(k)\displaystyle\mp\frac{\omega}{4\pi}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\pm\frac{i\omega}{4\pi}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\sinh\psi(k) (37)
=\displaystyle= ω2±iω2sinhψ=ω2±iλω2R02λ2.\displaystyle\mp\frac{\omega}{2}\pm i\frac{\omega}{2}\sinh\psi=\mp\frac{\omega}{2}\pm i\frac{\lambda\omega}{2\sqrt{R_{0}^{2}-\lambda^{2}}}.

In deriving Eq.(29), we used the property that ψ(k)\psi(k), defined by the relation tanhψ(k)=λ/R0{\rm tanh}\psi(k)=\lambda/R_{0}, is independent of kk and (dφ/dk)=1(d\varphi/dk)=1. Note that the behavior of the imaginary part of the quasi energies, predicted by the adiabatic analysis, well reproduces the exact curves, except near the phase transition point λ=λ¯c\lambda=\bar{\lambda}_{c} where the Berry phase contribution displays a singularity.

2. Second example. As a second example, let us consider the 𝒫𝒯\mathcal{PT}-symmetric two-level Hamiltonian

(k)=(iλt1+t2coskt1+t2coskiλ)\mathcal{H}(k)=\left(\begin{array}[]{cc}i\lambda&t_{1}+t_{2}\cos k\\ t_{1}+t_{2}\cos k&-i\lambda\end{array}\right) (38)

corresponding to G(k)=0G(k)=0, W(k)=1W(k)=1, R(k)=t1+t2coskR(k)=t_{1}+t_{2}\cos k, and φ(k)=0\varphi(k)=0, where t1t_{1} and t2t_{2} are real and positive parameters with t1>t2t_{1}>t_{2}. Since (dφ/dk)=0(d\varphi/dk)=0, the Berry phase vanishes and, according to the adiabatic analysis, when the system is slowly cycled with k=ωtk=\omega t the spectral phase transition of the quasi energies is sharp (exact) and occurs at the critical value λ¯c=t1t2\bar{\lambda}_{c}=t_{1}-t_{2} of the non-Hermitian parameter λ\lambda. The numerical computation of the quasi energies μ±\mu_{\pm} versus λ\lambda, as obtained by a direct numerical integration of the Schrödinger equation (15) using an accurate variable-step fourth-order Runge-Kutta method, confirms that the phase transition is sharp and the curves Im(μ±(λ)){\rm Im}(\mu_{\pm}(\lambda)) are well approximated by the behavior predicted by the adiabatic analysis, as shown in Fig.2.

Refer to caption
Figure 2: Behavior of the imaginary part of the quasi energies μ±\mu_{\pm} versus the non-Hermitian parameter λ\lambda for the cycled 𝒫𝒯\mathcal{PT}-symmetric system with Hamiltonian (k)\mathcal{H}(k) given by Eq.(30) for t1=1t_{1}=1, t2=0.5t_{2}=0.5 and for a modulation frequency (a) ω=0.02\omega=0.02, and (b) ω=0.1\omega=0.1. Open blue circles and red crosses refer to the exact curves, obtained from a numerical computation of quasi energies, and to the approximate curves obtained from the adiabatic analysis, respectivley. Note that the spectral phase transition is sharp around the critical point λ=λ¯c=t1t2=0.5\lambda=\bar{\lambda}_{c}=t_{1}-t_{2}=0.5.

3. Third example. As a third example, let us consider the 𝒫𝒯\mathcal{PT}-symmetric two-level Hamiltonian

(k)=(iλ+t0coskt1+t2exp(ik)t1+t2exp(ik)iλ+t0cosk)\mathcal{H}(k)=\left(\begin{array}[]{cc}i\lambda+t_{0}\cos k&t_{1}+t_{2}\exp(ik)\\ t_{1}+t_{2}\exp(-ik)&-i\lambda+t_{0}\cos k\end{array}\right) (39)

corresponding to G(k)=t0coskG(k)=t_{0}\cos k, W(k)=1W(k)=1, R(k)=t12+t22+2t1t2coskR(k)=\sqrt{t_{1}^{2}+t_{2}^{2}+2t_{1}t_{2}\cos k}, and φ(k)=atan[t2sink/(t1+t2cosk]\varphi(k)={\rm atan}[t_{2}\sin k/(t_{1}+t_{2}\cos k], where t0t_{0}, t1t_{1} and t2t_{2} are real and positive parameters with t1t2t_{1}\neq t_{2}. Since (dφ/dk)0(d\varphi/dk)\neq 0, the imaginary part of the Berry phase does not vanish and, according to the adiabatic analysis, when the system is slowly cycled with k=ωtk=\omega t the spectral phase transition of the quasi energies is imperfect (smooth). The phase transition occurs at the critical value λ¯c=|t2t1|\bar{\lambda}_{c}=|t_{2}-t_{1}| of the non-Hermitian parameter λ\lambda. The numerical computation of the quasi energies μ±\mu_{\pm} versus λ\lambda, as obtained by a direct numerical integration of the Schrödinger equation (15), confirms that the phase transition is imperfect and the curves Im(μ±(λ)){\rm Im}(\mu_{\pm}(\lambda)) are well approximated by the behavior predicted by the adiabatic analysis for λλ¯c\lambda\neq\bar{\lambda}_{c}, as shown in Fig.3. Note that in the slow-cycling regime [Fig.3(a)] the curves Im(μ±){\rm Im}(\mu_{\pm}) versus λ\lambda display a characteristic knee shape, indicating a smooth spectral phase transition. We remark that the terminology ”smooth” phase transition is meaningful in the adiabatic limit ω0\omega\rightarrow 0 solely, while when we cycle the system faster, so as ω\omega becomes comparable to the other characteristic frequencies of the Hamiltonian (such as the separation of adiabatic energies), the knee shape of the curves is continuously spoiled out and there is not any evident sharp transition of the imaginary part of the quasi energies as λ\lambda is increased; see Fig.3(b).

Refer to caption
Figure 3: Behavior of the imaginary part of the quasi energies μ±\mu_{\pm} versus the non-Hermitian parameter λ\lambda for the cycled 𝒫𝒯\mathcal{PT}-symmetric system with Hamiltonian (k)\mathcal{H}(k) given by Eq.(31) for t0=0.3t_{0}=0.3, t1=0.5t_{1}=0.5, t2=1t_{2}=1 and for a modulation frequency (a) ω=0.02\omega=0.02, and (b) ω=0.1\omega=0.1. Open blue circles and red crosses refer to the exact curves, obtained from a numerical computation of quasi energies, and to the approximate curves obtained from the adiabatic analysis. Note that in (a) (slow-cycling limit) the spectral phase transition is smooth around the critical point λ=λ¯c=|t2t1|=0.5\lambda=\bar{\lambda}_{c}=|t_{2}-t_{1}|=0.5, displaying a characteristic knee shape. In (b) the system is cycled faster and the knee shape of the curves is spoiled out. In both cases the adiabatic theory fails to the predict the correct behavior of the quasi energies near the critical point.

III Wannier-Stark ladder phase transition

The imperfect spectral phase transition, arising from the complex Berry phase in a slowly-cycled two-level system presented in the previous section, finds an interesting illustrative application to the problem of Wannier-Stark ladder formation in non-Hermitian lattices subjected to a weak external dc field and the transition from periodic to aperiodic Bloch-Zener oscillations recently observed for some models in Ref.b81 . In this case the Berry phase is also referred to as the Zak phase b17 , which is the geometric phase acquired during an adiabatic motion of a Bloch particle across the Brillouin zone.

III.1 Model

Let us consider a two-band tight-binding lattice model driven by a dc force FF. In physical space, the temporal evolution of the single-particle state of the system is described by the Schrödinger equation

idandt\displaystyle i\frac{da_{n}}{dt} =\displaystyle= lρnlal+lσnlblFnan\displaystyle\sum_{l}\rho_{n-l}a_{l}+\sum_{l}\sigma_{n-l}b_{l}-Fna_{n} (40)
idbndt\displaystyle i\frac{db_{n}}{dt} =\displaystyle= lθnlal+lηnlblFnbn\displaystyle\sum_{l}\theta_{n-l}a_{l}+\sum_{l}\eta_{n-l}b_{l}-Fnb_{n} (41)

for the amplitudes ana_{n} and bnb_{n} in the two sublattices A and B of the nn-th unit cell of the crystal. In the above equations, the coefficients ρ0\rho_{0} and η0\eta_{0} are the on-site energy potentials in the two sublattices A and B, respectively; ρl\rho_{l} and ηl\eta_{l} (l0l\neq 0) are the intra-dimer hopping amplitudes; finally, σl\sigma_{l} and θl\theta_{l} are the inter-dimer hopping amplitudes. In the absence of the dc force, i.e. for F=0F=0, we can assume an(t)=ψ1(t)exp(ikn)a_{n}(t)=\psi_{1}(t)\exp(ikn) and bn(t)=ψ2(t)exp(ikn)b_{n}(t)=\psi_{2}(t)\exp(ikn), where kk is the Bloch wave number that spans the Brillouin zone 0k2π0\leq k\leq 2\pi. In this case, from Eqs.(32) and (33) one obtains

iddt(ψ1ψ2)=(k)(ψ1ψ2)i\frac{d}{dt}\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\mathcal{H}(k)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right) (42)

where the elements of the 2×22\times 2 Bloch Hamiltonian (k)\mathcal{H}(k) are given by

11(k)=lρlexp(ikl)\displaystyle\mathcal{H}_{11}(k)=\sum_{l}\rho_{l}\exp(-ikl) (43)
12(k)=lσlexp(ikl)\displaystyle\mathcal{H}_{12}(k)=\sum_{l}\sigma_{l}\exp(-ikl) (44)
21(k)=lθlexp(ikl)\displaystyle\mathcal{H}_{21}(k)=\sum_{l}\theta_{l}\exp(-ikl) (45)
22(k)=lηlexp(ikl).\displaystyle\mathcal{H}_{22}(k)=\sum_{l}\eta_{l}\exp(-ikl). (46)

The Bloch Hamiltonian is 𝒫𝒯\mathcal{PT}-symmetric, with 𝒫=σx\mathcal{P}=\sigma_{x} and 𝒯=𝒦\mathcal{T}=\mathcal{K}, provided that

θl=σl,ηl=ρl.\theta^{*}_{-l}=\sigma_{l}\;,\;\;\eta_{-l}^{*}=\rho_{l}. (47)

Such conditions ensure that 22(k)=11(k)\mathcal{H}_{22}(k)=\mathcal{H}^{*}_{11}(k) and 21(k)=12(k)\mathcal{H}_{21}(k)=\mathcal{H}^{*}_{12}(k). In this case, the lattice does not display the non-Hermitian skin effect skin and the energy spectrum is absolutely continuous and composed by two energy bands, with the dispersion relation given by Eq.(6). A 𝒫𝒯\mathcal{PT} symmetry breaking phase transition of Bloch bands arises when the non-Hermitian parameter λ\lambda in the system is increased above the critical value λ¯c\bar{\lambda}_{c}, alike in the two-level system discussed in Sec.II.

III.2 Wannier-Stark ladders

When the external dc force is applied, i.e. for F0F\neq 0, the energy spectrum becomes pure point and composed by two Wannier-Stark ladders WS1 ; WS2 , with the allowed energies given by

El=lF±ΘE_{l}=lF\pm\Theta (48)

where l=0,±1,±2,±3,..l=0,\pm 1,\pm 2,\pm 3,.. and Θ\Theta describes the energy shift of the two ladders. The corresponding eigenstates are normalizable (localized) with a higher-than-exponential localization.
In an Hermitian lattice, the energy shift Θ\Theta is real and the dynamics in the time domain is generally aperiodic and corresponds to a superposition of Bloch oscillations and Zener tunneling between the two bands WS1 ; WS2 ; WS3 ; WS4 ; WS5 ; WS5bis ; WS6 ; WS7 . The dynamics is characterized by two time periods: The first one, T1=2π/FT_{1}=2\pi/F, is determined by the mode spacing of each WS ladder and is related to the Bloch oscillation dynamics, whereas the second one, T2=π/ΘT_{2}=\pi/\Theta, is determined by the shift of the two interleaved WS ladders.
In a non-Hermitian lattice the energy shift Θ\Theta can become complex and, as we show below, in the small-forcing limit F0F\rightarrow 0 it contains the complex Zak phase of the Bloch Hamiltonian (k)\mathcal{H}(k). More precisely, we will show below that Θ\Theta is the quasi energy μ+\mu_{+} of the Bloch Hamiltonian (k)\mathcal{H}(k), cycled over the Brillouin zone at a frequency ω=F\omega=F, i.e. with k=Ftk=Ft. This means that the WS energy spectrum undergoes a phase transition as λ\lambda is increased above λ¯c\bar{\lambda}_{c}, from real to complex energies, and the phase transition can be either sharp or smooth, depending on whether the imaginary part of the Zak phase for λ<λ¯c\lambda<\bar{\lambda}_{c} is vanishing or not.
To calculate the WS energy spectrum EE, let us assume an(t)=a¯nexp(iEt)a_{n}(t)=\bar{a}_{n}\exp(-iEt), bn(t)=b¯nexp(iEt)b_{n}(t)=\bar{b}_{n}\exp(-iEt) in Eqs.(32-33), and let us introduce the spectral variables

ψ1(k)\displaystyle\psi_{1}(k) =\displaystyle= exp(iEk/F)na¯nexp(ikn)\displaystyle\exp(-iEk/F)\sum_{n}\bar{a}_{n}\exp(-ikn) (49)
ψ2(k)\displaystyle\psi_{2}(k) =\displaystyle= exp(iEk/F)nb¯nexp(ikn).\displaystyle\exp(-iEk/F)\sum_{n}\bar{b}_{n}\exp(-ikn). (50)

It readily follows that ψ1,2(k)\psi_{1,2}(k) satisfy the Sturm-Liouville problem

iFddk(ψ1ψ2)=(k)(ψ1ψ2)iF\frac{d}{dk}\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\mathcal{H}(k)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right) (51)

on the interval 0k2π0\leq k\leq 2\pi, with the boundary conditions

ψ1,2(2π)=ψ1,2(0)exp(2πiEF).\psi_{1,2}(2\pi)=\psi_{1,2}(0)\exp\left(-\frac{2\pi iE}{F}\right). (52)

Once the spectral amplitudes ψ1,2(k)\psi_{1,2}(k) and eigenenergies EE have been determined, the eigenvectors (a¯n,b¯n)(\bar{a}_{n},\bar{b}_{n}), corresponding to the energy EE, are determined using the inverse relations

a¯n\displaystyle\bar{a}_{n} =\displaystyle= 12π02π𝑑kψ1(k)exp(ikn+iEk/F)\displaystyle\frac{1}{2\pi}\int_{0}^{2\pi}dk\psi_{1}(k)\exp(ikn+iEk/F) (53)
b¯n\displaystyle\bar{b}_{n} =\displaystyle= 12π02π𝑑kψ2(k)exp(ikn+iEk/F).\displaystyle\frac{1}{2\pi}\int_{0}^{2\pi}dk\psi_{2}(k)\exp(ikn+iEk/F). (54)

Interestingly, after letting k=ωtk=\omega t Eq.(43) indicates that ψ1,2(k)\psi_{1,2}(k) can be viewed as the amplitudes of a two-level 𝒫𝒯\mathcal{PT}-symmetric system, with Hamiltonian (k)\mathcal{H}(k), which is slowly cycled in time at the frequency ω=F\omega=F. This basically corresponds to the fact that in Bloch space the external force introduces a uniform drift of the quasi-momentum kk to span the entire Brillouin zone. The Sturm-Liouville problem, defined by Eqs.(43) and (44), can be solved as follows. Let us indicate by 𝝍+\bm{\psi}_{+} and 𝝍\bm{\psi}_{-} the eigenvectors of the Floquet matrix \mathcal{R}, introduced in Sec.II.B, with eigenvalues (quasi energies) μ±\mu_{\pm}. Then Eq.(43) is satisfied by letting either

(ψ1(k)ψ2(k))=𝒰(kF)exp(ik/F)𝝍+\displaystyle\left(\begin{array}[]{c}\psi_{1}(k)\\ \psi_{2}(k)\end{array}\right)=\mathcal{U}\left(\frac{k}{F}\right)\exp(-i\mathcal{R}k/F)\bm{\psi}_{+} (57)
=exp(iμ+k/F)𝒰(kR)𝝍+\displaystyle=\exp(-i\mu_{+}k/F)\mathcal{U}\left(\frac{k}{R}\right)\bm{\psi}_{+}

or

(ψ1(k)ψ2(k))=𝒰(kF)exp(ik/F)𝝍\displaystyle\left(\begin{array}[]{c}\psi_{1}(k)\\ \psi_{2}(k)\end{array}\right)=\mathcal{U}\left(\frac{k}{F}\right)\exp(-i\mathcal{R}k/F)\bm{\psi}_{-} (60)
=exp(iμk/F)𝒰(kF)𝝍.\displaystyle=\exp(-i\mu_{-}k/F)\mathcal{U}\left(\frac{k}{F}\right)\bm{\psi}_{-}.

Since 𝒰(0)=𝒰(2π/F)=\mathcal{U}(0)=\mathcal{U}(2\pi/F)=\mathcal{I} (the 2×22\times 2 identity matrix), to satisfy the boundary conditions Eq.(44) one should have

2πFμ±=2πEF2lπ\frac{2\pi}{F}\mu_{\pm}=\frac{2\pi E}{F}-2l\pi

i.e.

E=lF+μ±E=lF+\mu_{\pm} (61)

where l=0,±1,±2,l=0,\pm 1,\pm 2,.... Equation (47) provides the general form of the Wannier-Stark ladders of allowed energies in terms of the quasi energies μ±\mu_{\pm} of the cycled two-level Bloch Hamiltonian (k)\mathcal{H}(k). It is precisely Eq.(40) with the energy shift parameter Θ\Theta given by Θ=μ+\Theta=\mu_{+}.
When the external dc force is weak, i.e. in the limit F0F\rightarrow 0, the Bloch wave number k=Ftk=Ft in the two-level Bloch Hamiltonian (k)\mathcal{H}(k) varies slowly with time, and thus the quasi energies can be approximated by Eq.(23). One obtains

E=lF+12π02π𝑑kE±(k)+F2πγB±E=lF+\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{\pm}(k)+\frac{F}{2\pi}\gamma_{B_{\pm}} (62)

corresponding to the energy shift

Θ=μ+=12π02π𝑑kE+(k)+F2πγB+.\Theta=\mu_{+}=\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{+}(k)+\frac{F}{2\pi}\gamma_{B_{+}}. (63)
Refer to caption
Figure 4: Schematic of two binary non-Hermitian lattices displaying (a) perfect, and (b) imperfect Wannier-Stark phase transitions under a dc force. t0t_{0}, t1t_{1} and t2t_{2} are Hermitian hopping amplitudes. The non-Hermiticity in the system is provided by the gain and loss terms ±λ\pm\lambda in the two sublattices A and B. The Bloch Hamiltonian of the two lattices is given by Eq.(30) for model (a), and by Eq.(31) for model (b).

Therefore, under a weak external driving the Wannier-Stark energy spectrum undergoes a phase transition at λ=λ¯c\lambda=\bar{\lambda}_{c}, which is either sharp or smooth depending on whether the imaginary part of the Zak phase γB+\gamma_{B{{}_{+}}} is vanishing or not when λ<λ¯c\lambda<\bar{\lambda}_{c}. We mention that, contrary to other non-Hermitian lattice models where the spectral (𝒫𝒯\mathcal{PT} symmetry breaking) phase transition coincides with a localization/delocalization phase transition uffa1 ; uffa1bis ; uffa2 , in the Wannier-Stark ladder problem the spectral phase transition does not correspond to a localization/delocalization phase transition because the eigenstates of the Wannier-Stark Hamiltonian are always localized, for both λ<λ¯c\lambda<\bar{\lambda}_{c} and λλ¯c\lambda\geq\bar{\lambda}_{c}. This very general result follows from the fact that the spectral amplitudes ψ1,2(k)exp(iEk/F)\psi_{1,2}(k)\exp(iEk/F), with ψ1,2(k)\psi_{1,2}(k) solutions to the Sturm-Liouville problem [Eqs.(43,44)], are periodic and continuously differentiable functions of kk and thus their Fourier coefficients a¯n\bar{a}_{n}, b¯n\bar{b}_{n} decay as n±n\rightarrow\pm\infty at least like 1/n\sim 1/n, regardless of the value of λ\lambda.
In a non-Hermitian lattice below the Wannier-Stark phase transition point, i.e. for λ<λ¯c\lambda<\bar{\lambda}_{c}, the dynamical signature of a complex Zak phase can be probed looking at the temporal behavior of Bloch-Zener oscillations b81 . When the imaginary part of the Zak phase is vanishing, the temporal dynamics is rather generally aperiodic and characterized by the two periods T1T_{1} and T2T_{2}, like in an ordinary Hermitian lattice under a dc field: only accidentally the dynamics can be periodic. On the other hand, when the imaginary part of the complex Zak phase does not vanish, after an initial transient the dynamics becomes periodic with period T1T_{1}. In fact, a rather arbitrary excitation of the system at initial time t=0t=0 can be decomposed as a superposition of localized Wannier-Stark eigenstates belonging to the two ladders, and the dynamics at successive times is governed by the interference of such localized eigenstates. The localized Wannier-Stark eigenstates in one ladder, excited by the initial condition, decay in time with a damping rate FIm(γB+)\sim F{\rm{Im}}(\gamma_{B_{+}}), while the eigenstates in the other ladder are amplified in time with an amplification rate FIm(γB+)\sim F{\rm{Im}}(\gamma_{B_{+}}). Therefore, after a transient time of order 1/FIm(γB+)\sim 1/F{\rm{Im}}(\gamma_{B_{+}}) only the Wannier-Stark eigenstates in the former ladder survive and the dynamics become periodic with the period T1T_{1} b81 .

As illustrative examples, let us consider the binary lattices depicted in Figs.4(a) and 4(b). The non-Hermiticity in the lattices is introduced by assuming energy gain and loss terms ±λ\pm\lambda in the two sub lattices A and B. The binary lattice of Fig.4(a) was introduced in a previous work b74 and its Bloch Hamiltonian (k)\mathcal{H}(k) is given by Eq.(30), previously introduced in Sec.II.C. Since (dφ/dk)0(d\varphi/dk)\equiv 0, the Wannier-Stark ladder phase transition in this model is sharp. The model shown in Fig.4(b) is a non-Hermitian extension of the Rice-Mele model b54basta1 and its Bloch Hamiltonian is given by Eq.(31). For this model, the spectral phase transition of the Wannier-Stark energies is imperfect. We emphasize that our analysis is very general and could be applied to a generic 𝒫𝒯\mathcal{PT}-symmetric binary lattice, also displaying long-range hopping.

IV Conclusions and discussion

In many classical and quantum systems described by an effective non-Hermitian Hamiltonian, where energy and particles can be exchanged with external reservoirs, the energies of the Hamiltonian are rather generally complex. However, in certain classes of non-Hermitian systems the energy spectrum can remain entirely real in spite of non-Hermiticity. A paradigmatic example is provided by systems possessing parity-time symmetry, where the energy spectrum remains entirely real in the unbroken 𝒫𝒯\mathcal{PT} phase. When the strength of non-Hermiticity in the system is increased, a spectral phase transition to complex energies is usually observed, corresponding to the unbroken 𝒫𝒯\mathcal{PT} phase. Such spectral phase transitions are universally sharp. In this work we considered periodically and slowly cycled non-Hermitian models possessing instantaneous 𝒫𝒯\mathcal{PT} symmetry and showed that the phase transition can remain exact (sharp) or become imperfect (smooth) when the strength of non-Hermiticity in the system is increased above a critical value. The imperfect nature of the phase transition in the latter case is universally ascribable to a non-vanishing imaginary part of the complex Berry phase associated to the cyclic adiabatic evolution of the system. This remarkable phenomenon has been illustrated by considering a rather general class of 𝒫𝒯\mathcal{PT}-symmetric two-level systems, for which a rigorous adiabatic analysis both below and above the phase transition point has been developed. The results have been applied to describe the spectral phase transitions of the Wannier-Stark ladders in a broad class of 𝒫𝒯\mathcal{PT}-symmetric two-band non-Hermitian lattices subjected to an external dc field, however our analysis is expected to hold for more general multi-band systems. In fact, under the adiabatic conditions and assuming no state flip after one adiabatic cycle, the form of quasi energies can be given in terms of dynamic and geometric (Zak) phases, and the complex or real nature of the latter defines the smooth or sharp nature of the spectral phase transitions in the slow-cycling regime. Our results provide fresh and novel insights into phase transitions of open quantum or classical systems, providing important examples of smooth phase transitions in non-Hermitian physics and unraveling the main role played by the non-Hermitian Berry phase.

Acknowledgements.
S.L. acknowledges the Spanish State Research Agency, through the Severo Ochoa and Maria de Maeztu Program for Centers and Units of Excellence in R&D (Grant No. MDM-2017-0711). L.F. acknowledges the support from National Science Foundation (NSF) (ECCS-1846766).

Appendix A Berry connection and Berry phase

For the cycled two-level 𝒫𝒯\mathcal{PT} symmetric model considered in Sec.II.A, the elements of the 2×22\times 2 matrix of the non-Hermitian Berry connection are given in terms of the biorthogonal product as

𝒜n,l=i𝐯n|k𝐮l\mathcal{A}_{n,l}=-i\langle\mathbf{v}_{n}|\partial_{k}\mathbf{u}_{l}\rangle (64)

where n,ln,l take the values ++ or -. Using Eqs.(7,8) and (10,11) given in the main text, the explicit form of the Berry connection can be readily calculated and read

𝒜+,+\displaystyle\mathcal{A}_{+,+} =\displaystyle= dφdksin2(θ2)\displaystyle-\frac{d\varphi}{dk}\sin^{2}\left(\frac{\theta}{2}\right) (65)
𝒜+,\displaystyle\mathcal{A}_{+,-} =\displaystyle= 12idθdk+12dφdksinθ\displaystyle-\frac{1}{2}i\frac{d\theta}{dk}+\frac{1}{2}\frac{d\varphi}{dk}\sin\theta (66)
𝒜,+\displaystyle\mathcal{A}_{-,+} =\displaystyle= 12idθdk+12dφdksinθ\displaystyle\frac{1}{2}i\frac{d\theta}{dk}+\frac{1}{2}\frac{d\varphi}{dk}\sin\theta (67)
𝒜,\displaystyle\mathcal{A}_{-,-} =\displaystyle= dφdkcos2(θ2).\displaystyle-\frac{d\varphi}{dk}\cos^{2}\left(\frac{\theta}{2}\right). (68)

The Berry phases associated to the two adiabatically-evolving eigenstates 𝐮±(k)\mathbf{u}_{\pm}(k) are given by

γB+\displaystyle\gamma_{B_{+}} \displaystyle\equiv 02π𝑑k𝒜+,+=02π𝑑kdφdksin2(θ2)\displaystyle\int_{0}^{2\pi}dk\mathcal{A}_{+,+}=-\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\sin^{2}\left(\frac{\theta}{2}\right)\;\;\;\; (69)
γB\displaystyle\gamma_{B_{-}} \displaystyle\equiv 02π𝑑k𝒜,=02π𝑑kdφdkcos2(θ2).\displaystyle\int_{0}^{2\pi}dk\mathcal{A}_{-,-}=-\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\cos^{2}\left(\frac{\theta}{2}\right).\;\;\;\; (70)

From Eqs.(A6) and (A7) it readily follows that γB++γB=02π𝑑k(dφ/dk)\gamma_{B_{+}}+\gamma_{B_{-}}=-\int_{0}^{2\pi}dk(d\varphi/dk) is either zero or an integer multiple than 2π2\pi. Since the Berry phase is defined apart from integer multiples than 2π2\pi, we can thus write

γB+\displaystyle\gamma_{B_{+}} =\displaystyle= γB=\displaystyle-\gamma_{B_{-}}=
=\displaystyle= 1202π𝑑kdφdk+1202π𝑑kcosθdφdk.\displaystyle-\frac{1}{2}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}+\frac{1}{2}\int_{0}^{2\pi}dk\cos\theta\frac{d\varphi}{dk}.

The complex angle θ=θ(k)\theta=\theta(k) is defined by Eq.(9) given in the main text, i.e.

tanθ(k)=R(k)iλW(k)\tan\theta(k)=\frac{R(k)}{i\lambda W(k)} (72)

which can be solved by letting

θ(k)=π/2iψ(k),\theta(k)=\pi/2-i\psi(k), (73)

where the function ψ(k)\psi(k) is given by

tanhψ(k)=λW(k)R(k).\tanh\psi(k)=\frac{\lambda W(k)}{R(k)}. (74)

Using Eqs.(A8) and (A10), one finally obtains

γB±=1202π𝑑kdφdk±i202π𝑑kdφdksinhψ(k)\gamma_{B_{\pm}}=\mp\frac{1}{2}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\pm\frac{i}{2}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\sinh\psi(k) (75)

The above expression of the Berry phase is formally valid for any value of the non-Hermitian parameter, except for λ=λ¯c\lambda=\bar{\lambda}_{c}. Note that for λ<λ¯c\lambda<\bar{\lambda}_{c} one has |λW(k)/R(k)|<1|\lambda W(k)/R(k)|<1 and thus the function ψ(k)\psi(k) is real over the entire interval 0k2π0\leq k\leq 2\pi. In this case Eq.(A12) shows that the real part of the Berry phase is quantized and can take only the two values 0 or π\pi (mod. 2π2\pi), whereas the imaginary part of the Berry phase is not quantized and vanishes whenever (dφ/dk)0(d\varphi/dk)\equiv 0.

Appendix B Adiabatic analysis

In this Appendix we derive an analytical expression of the quasi energies μ±\mu_{\pm} of the cycled two-level 𝒫𝒯\mathcal{PT} symmetric system, considered in Sec.II.B of the main text, in the adiabatic limit of slow cycling ω0\omega\rightarrow 0. It should be mentioned that special attention is required when using adiabatic methods to slowly-evolving non-Hermitian systems because: (i) Owing to possible nontrivial topologies of the energy curves in complex plane, even in absence of eigenvalue degeneracies it could happen that an adiabatically evolving eigenstate, after one cycle, does not come back to its initial state because of energy and eigenvector flipping b58 ; b59 ; (ii) Even in the slow cycling regime the adiabatic approximation can easily break down when the instantaneous eigenenergies are complex bf1 ; bf2 ; bf3 , and the adiabatic approximation can be safely applied only to the most dominant eigenstate of the system.
After letting k=ωtk=\omega t, the Schrödinger equation (15) reads

iωddk(ψ1ψ2)=(k)(ψ1ψ2).i\omega\frac{d}{dk}\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)=\mathcal{H}(k)\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right). (76)

To perform the adiabatic analysis, let us distinguish two cases.
First case: λ<λ¯c\lambda<\bar{\lambda}_{c}. In this case the two energy curves E±(k)E_{\pm}(k), as kk spans the interval (0,2π)(0,2\pi), are straight and non-intersecting segments on the real energy axis; see Fig.5(a). Therefore, the energy curves are line gapped and there is not any eigenvalue/eigenstate flip after one cycle. From the point of view of the adiabatic analysis, the system thus behaves like an Hermitian one, even though the Hamiltonian is not Hermitian. We then expand the state vector (ψ1(k),ψ2(k))T(\psi_{1}(k),\psi_{2}(k))^{T} as a superposition of the instantaneous eigenstates 𝐮+(k)\mathbf{u}_{+}(k) and 𝐮(k)\mathbf{u}_{-}(k) of (k)\mathcal{H}(k), i.e. let us set

(ψ1(k)ψ2(k))\displaystyle\left(\begin{array}[]{c}\psi_{1}(k)\\ \psi_{2}(k)\end{array}\right) =\displaystyle= a+(k)𝐮+(k)exp{iω0k𝑑ξE+(ξ)}\displaystyle a_{+}(k)\mathbf{u}_{+}(k)\exp\left\{-\frac{i}{\omega}\int_{0}^{k}d\xi E_{+}(\xi)\right\} (79)
+\displaystyle+ a(k)𝐮(k)exp{iω0k𝑑ξE(ξ)}\displaystyle a_{-}(k)\mathbf{u}_{-}(k)\exp\left\{-\frac{i}{\omega}\int_{0}^{k}d\xi E_{-}(\xi)\right\}\;\;\;\;\;\; (80)

where a+(k)a_{+}(k) and a(k)a_{-}(k) are the adiabatic amplitudes and

E±(k)=±R2(k)λ2W2(k)E_{\pm}(k)=\pm\sqrt{R^{2}(k)-\lambda^{2}W^{2}(k)} (81)

are the instantaneous eigenenergies. The evolution equations of the amplitudes a±(k)a_{\pm}(k) are readily obtained by substitution of the Ansatz (B2) into Eq.(B1) and taking the scalar product of the equation so obtained by 𝐯+|\langle\mathbf{v}_{+}| and 𝐯|\langle\mathbf{v}_{-}|. Using the biorthogonal conditions (12), one obtains

ida+dk\displaystyle i\frac{da_{+}}{dk} =\displaystyle= 𝒜+,+a++\displaystyle\mathcal{A}_{+,+}a_{+}+
+\displaystyle+ 𝒜+,aexp{iω0k𝑑ξ[E+(ξ)E(ξ)]}\displaystyle\mathcal{A}_{+,-}a_{-}\exp\left\{\frac{i}{\omega}\int_{0}^{k}d\xi\left[E_{+}(\xi)-E_{-}(\xi)\right]\right\}
idadk\displaystyle i\frac{da_{-}}{dk} =\displaystyle= 𝒜,a+\displaystyle\mathcal{A}_{-,-}a_{-}+
+\displaystyle+ 𝒜,+a+exp{iω0k𝑑ξ[E+(ξ)E(ξ)]}\displaystyle\mathcal{A}_{-,+}a_{+}\exp\left\{-\frac{i}{\omega}\int_{0}^{k}d\xi\left[E_{+}(\xi)-E_{-}(\xi)\right]\right\}

where 𝒜n,l\mathcal{A}_{n,l} (n,l=+,)n,l=+,-) is the non-Hermitian Berry connection, given by Eqs.(A1-A5). In the adiabatic limit ω0\omega\rightarrow 0, since the energy difference E+(k)E(k)E_{+}(k)-E_{-}(k) is entirely real and non-vanishing over the interval 0k2π0\leq k\leq 2\pi, the rapidly oscillating terms on the right hand sides of Eqs.(B4) and (B5) do not induce on average transitions between the two adiabatic amplitudes and can be disregarded (rotating-wave approximation). Hence one obtains

a±(2π)a±(0)exp(iγB±)a_{\pm}(2\pi)\simeq a_{\pm}(0)\exp(-i\gamma_{B_{\pm}}) (84)

where γB±\gamma_{B_{\pm}} are the Berry phases associated to the two adiabatically-evolving eigenstates 𝐮±(k)\mathbf{u}_{\pm}(k). The explicit form of the Berry phases is given by Eq.(A12).
The quasi energies μ±\mu_{\pm} are the eigenvalues of the matrix \mathcal{R}, which is obtained from the condition [Eq.(17) in the main text with t=2π/ωt=2\pi/\omega]

(ψ1(2π/ω)ψ2(2π/ω))=exp(2πi/ω)(ψ1(0)ψ2(0)).\left(\begin{array}[]{c}\psi_{1}(2\pi/\omega)\\ \psi_{2}(2\pi/\omega)\end{array}\right)=\exp(-2\pi i\mathcal{R}/\omega)\left(\begin{array}[]{c}\psi_{1}(0)\\ \psi_{2}(0)\end{array}\right). (85)

From Eqs.(B2) and (B5) it readily follows that 𝐮±(0)\mathbf{u}_{\pm}(0) are the eigenvectors of \mathcal{R} with corresponding quasi energies given by

μ±=12π02π𝑑kE±(k)+ω2πγB±.\mu_{\pm}=\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{\pm}(k)+\frac{\omega}{2\pi}\gamma_{B_{\pm}}. (86)

Note that, since E(k)=E+(k)E_{-}(k)=-E_{+}(k) and γB=γB+\gamma_{B_{-}}=-\gamma_{B_{+}}, the two quasi energies are opposite one another, i.e. μ=μ+\mu_{-}=-\mu_{+}, as it should be whenever G(k)=0G(k)=0. Note also that each quasi energy is given by the sum of two terms. The first term on the right hand side of Eq.(B8) is the usual dynamical phase term that one would obtain by a standard WKB analysis neglecting the Berry phase b77 ; b81 , whereas the second term on the right hand side of Eq.(B8) is the Berry phase contribution. While the dynamical phase term is always real and independent of the modulation frequency ω\omega, the Berry phase contribution vanishes as ω0\omega\rightarrow 0 and can display a nonvanishing imaginary part in models where (dφ/dk)0(d\varphi/dk)\neq 0. Therefore, we may conclude that for λ<λ¯c\lambda<\bar{\lambda}_{c} the imaginary part of the quasi energies, as predicted by the adiabatic analysis, reads

Im(μ±)=ω2πIm(γB±)=±ω4π02π𝑑kdφdksinhψ(k){\rm\text{Im}\,}(\mu_{\pm})=\frac{\omega}{2\pi}{\rm{Im}}(\gamma_{B_{\pm}})=\pm\frac{\omega}{4\pi}\int_{0}^{2\pi}dk\frac{d\varphi}{dk}\sinh\psi(k) (87)

where the real function ψ(k)\psi(k) is defined by Eq.(A11).

Refer to caption
Figure 5: Schematic behavior of the energy curves E±(k)E_{\pm}(k) of the 𝒫𝒯\mathcal{PT}-symmetric two-level Hamiltonian (k)\mathcal{H}(k) in complex energy plane as kk spans the interval 0k2π0\leq k\leq 2\pi (solid lines). In (a) λ<λ¯c\lambda<\bar{\lambda}_{c}, the two energy curves lie on the real energy axis and are line gapped. In (b) λ>λ¯c\lambda>\bar{\lambda}_{c} and the two energy curves cross at E=0E=0 (instantaneous exceptional point) at the critical values k=kck=k_{c} such that λW(kc)=±R(kc)\lambda W(k_{c})=\pm R(k_{c}). The dashed curves in (b) show the behavior of the energy curves for the modified Hamiltonian ϵ(k)\mathcal{H}_{\epsilon}(k), which avoids energy crossing and exceptional points.

Second case: λ>λ¯c\lambda>\bar{\lambda}_{c}. In this case the two energy curves E±(k)E_{\pm}(k), as kk spans the interval (0,2π)(0,2\pi), may touch one another at E=0E=0, as shown by the solid curves in Fig.5(b). The crossing occurs when kk equals the critical values k=kck=k_{c} such that λW(kc)=±R(kc)\lambda W(k_{c})=\pm R(k_{c}). At such points, the instantaneous Hamiltonian (kc)\mathcal{H}(k_{c}) is not diagonalizable and displays an exceptional point. Eventually, if W(k)W(k) does not vanish in the entire range (0,2π)(0,2\pi), at large values of λ\lambda the two energy curves can become separated and fully lie on the imaginary axis.
The occurrence of the instantaneous exceptional points and energy curve touching at k=kck=k_{c} during the cycle when λ>λ¯c\lambda>\bar{\lambda}_{c} makes it formally invalid the adiabatic analysis discussed in the previous case. To overcome such a limitation, we slightly modify the Hamiltonian of the system, from (k)\mathcal{H}(k) to ϵ(k)\mathcal{H}_{\epsilon}(k), by letting

(ϵ(k))11=iλW(k)+ϵ,(ϵ(k))22=(ϵ(k))11\left(\mathcal{H}_{\epsilon}(k)\right)_{11}=i\lambda W(k)+\epsilon\;,\;\;\left(\mathcal{H}_{\epsilon}(k)\right)_{22}=-\left(\mathcal{H}_{\epsilon}(k)\right)_{11} (88)

where ϵ>0\epsilon>0 is a small real parameter. For ϵ=0\epsilon=0 we recover the original Hamiltonian (k)\mathcal{H}(k). The instantaneous eigenenergies of ϵ(k)\mathcal{H}_{\epsilon}(k) read

Eϵ±=±R2(k)(λW(k)iϵ)2.E_{\epsilon\;\pm}=\pm\sqrt{R^{2}(k)-(\lambda W(k)-i\epsilon)^{2}}. (89)

A non-vanishing (albeit small) value of ϵ\epsilon breaks exact 𝒫𝒯\mathcal{PT} symmetry and avoids the energy curve touching and the appearance of the instantaneous exceptional points during the adiabatic cycle, as shown by the dashed curves in Fig.5(b). Since the two energy curves are now line gapped and there are not exceptional points along the cycle, we can again expand the state vector of the system as a superposition of the instantaneous eigenstates of ϵ(k)\mathcal{H}_{\epsilon}(k) with adiabatic amplitudes a±(k)a_{\pm}(k), which evolve according to Eqs.(B4) and (B5) (these are exact equations). The main difference is that the Berry connection and instantaneous eigenenergies entering in such equations are now those of the modified Hamiltonian ϵ(k)\mathcal{H}_{\epsilon}(k) rather than (k)\mathcal{H}(k). The expressions of the instantaneous (right and left) eigenstates of ϵ(k)\mathcal{H}_{\epsilon}(k), and thus of Berry connection and Berry phases, are formally the same as those of (k)\mathcal{H}(k), with the complex angle θ=θ(k)\theta=\theta(k) now defined by the relation

tanθ(k)=R(k)iλW(k)+ϵ.\tan\theta(k)=\frac{R(k)}{i\lambda W(k)+\epsilon}. (90)

It should be noted that, as ϵ0\epsilon\rightarrow 0 the imaginary part of θ(k)\theta(k) diverges at the critical values k=kck=k_{c}, however for a linear crossing, such that λW(kc)±R(kc)\lambda W^{\prime}(k_{c})\neq\pm R^{\prime}(k_{c}), the singularity of cosθ(k)\cos\theta(k) near k=kck=k_{c} is of the type cosθ(k)1/kkc\cos\theta(k)\sim 1/\sqrt{k-k_{c}} and thus integrable, leading to a finite value of the Berry phase according to Eq.(A8).
To calculate the quasi energies, we exploit the fact that μ=μ+\mu_{-}=-\mu_{+}, so that we can compute the quasi energy of the dominant adiabatic eigenstate of the system, i.e. with the largest imaginary part of instantaneous energy [corresponding to the dashed curve in the first quadrant of Fig.5(b)]. For such a state we can in fact safely apply the adiabatic approximation, avoiding the problem of adiabaticity breakdown that could arise for the non-dominant eigenstate bf1 ; bf2 . For example, assuming that 𝐮+(k)\mathbf{u}_{+}(k) is the dominant instantaneous eigenstate, i.e. with Im(E+(k))0{\rm Im}(E_{+}(k))\geq 0, we can safely apply the rotating-wave approximation to the second term on the right hand side of Eq.(B4), thus obtaining a+(2π)a+(0)exp(iγB+)a_{+}(2\pi)\simeq a_{+}(0)\exp(-i\gamma_{B_{+}}). Proceeding as in the previous case, in the adiabatic limit one then obtains the following expression of the quasi energy μ+\mu_{+}, associated to the dominant adiabatic eigenstate

μ+=12π02π𝑑kE+(k)+ω2πγB+.\mu_{+}=\frac{1}{2\pi}\int_{0}^{2\pi}dkE_{+}(k)+\frac{\omega}{2\pi}\gamma_{B_{+}}. (91)

The other quasi energy is then given by μ=μ+\mu_{-}=-\mu_{+}. This result formally coincides with the one obtained in the case λ<λ¯c\lambda<\bar{\lambda}_{c} [see Eq.(B8)], however in Eq.(B13) the Berry phase term should be obtained from the modified angle θ\theta, given by Eq.(B12), and then taking the limit ϵ0\epsilon\rightarrow 0. The main difference in the λ>λ¯c\lambda>\bar{\lambda}_{c} case is that the dynamical phase contribution to the quasi energy has a non-vanishing imaginary part, which dominates over the imaginary contribution of the Berry phase term in the adiabatic (ω0\omega\rightarrow 0) limit.
Finally, we mention that at the phase transition point λ=λ¯c\lambda=\bar{\lambda}_{c} the crossing of the exceptional point during the oscillation cycle, at k=kck=k_{c}, is quadratic rather than linear, i.e. one has λW(kc)=±R(kc)\lambda W(k_{c})=\pm R(k_{c}) and λW(kc)=±R(kc)\lambda W{{}^{\prime}}(k_{c})=\pm R{{}^{\prime}}(k_{c}). In this case, as ϵ0\epsilon\rightarrow 0 the singularity of cosθ(k)\cos\theta(k) near k=kck=k_{c} is of the type cosθ(k)1/(kkc)\cos\theta(k)\sim 1/(k-k_{c}) and thus it is not integrable, leading to a diverging value of the Berry phase according to Eq.(A8). Therefore, the adiabatic analysis fails to predict the correct values of the quasi energies as λ\lambda approaches the critical value λ¯c\bar{\lambda}_{c}, either from below or from above. Such a failure is clearly illustrated in the exactly-solvable model with Hamiltonian (k)\mathcal{H}(k) given by Eq.(24), discussed in the main text [see specifically Eq.(29) and Fig.1(b)].

References

  • (1) M.V. Berry, Quantal phase factors accompanying adiabatic changes, Proc. R. Soc. A 392, 45 (1984).
  • (2) A. Shapere and F. Wilczek, Geometric Phases in Physics (World Scientific, Singapore, 1989).
  • (3) A. Bohm, A. Mostafazadeh, H. Koizumi, Q. Niu, and J. Zwanziger, The Geometrie Phase in Quantum Systems (Springer, Berlin, 2003).
  • (4) E. Cohen, H. Larocque, F. Bouchard, F. Nejadsattari, Y. Gefen, and E. Karimi, Geometric phase from Aharonov-Bohm to Pancharatnam-Berry and beyond, Nat. Rev. Phys. 1, 437 (2019).
  • (5) D. Ellinas, S. M. Barnett, and M. A. Dupertuis, Berry phase in optical resonance, Phys. Rev. A 39, 3228 (1989).
  • (6) E.R. Meyer, A.E. Leanhardt, E.A. Cornell, and J.L. Bohn, Berry-like phases in structured atoms and molecules, Phys. Rev. A 80, 062110 (2009).
  • (7) P.V. Mironova, M.A. Efremov, and W.P. Schleich, Berry phase in atom optics, Phys. Rev. A 87, 013627 (2013).
  • (8) R. Resta, Manifestations of Berry phase in molecules and condensed matter, J. Phys.: Condens. Matter 12, R107 (2000).
  • (9) D. Xiao, M.-C. Chang, and Q. Niu. Berry phase effects on electronic properties, Rev. Mod. Phys. 82, 1959 (2010).
  • (10) J. Cayssol and J.N. Fuchs, Topological and geometrical aspects of band theory, J. Phys. Mater 4, 034007 (2021).
  • (11) R. Simon, H.J. Kimble, and E.C.G. Sudarshan, Evolving geometric phase and its dynamical manifestation as a frequency shift: an optical experiment, Phys. Rev. Lett. 61, 19 (1988).
  • (12) R.Y. Chiao, Berry Phases in Optics, In: van Haeringen, W., Lenstra, D. (eds), Analogies in Optics and Micro Electronics (Springer, Dordrecht, 1990).
  • (13) C. P. Jisha, S. Nolte, and A. Alberucci, Geometric Phase in Optics: From Wavefront Manipulation to Waveguiding, Laser & Photon. Rev. 15, 2100003 (2021).
  • (14) K.G. Wilson, Confinement of quarks, Phys. Rev. D 10, 2445 (1974).
  • (15) M. Stone, V. Dwivedi, and T. Zhou, Berry phase, Lorentz covariance, and anomalous velocity for Dirac and Weyl particles, Phys. Rev. D 91, 025004 (2015).
  • (16) M. Baggio, V. Niarchos, and K. Papadodimas, Aspects of Berry phase in QFT, J. High Energy Phys. 2017, 62 (2017).
  • (17) B.K. Pal, S. Pal, and B, Basu, The Berry phase in inflationary cosmology, Class. Quantum Grav. 30, 125002 (2013).
  • (18) J. Zak, Berry phase for energy bands in solids, Phys Rev Lett 62, 2747 (1989).
  • (19) M. Atala, M. Aidelsburger, J. T. Barreiro, D. Abanin, T. Kitagawa, E. Demler, and I. Bloch, Direct measurement of the Zak phase in topological Bloch bands, Nat. Phys. 9, 795 (2013).
  • (20) M. Xiao, Z.Q. Zhang, and C.T. Chan, Surface Impedance and Bulk Band Geometric Phases in One-Dimensional Systems, Phys. Rev. X 4, 021017 (2014).
  • (21) Q. Wang, M. Xiao, H. Liu, S. Zhu, and C. T. Chan, Measurement of the Zak phase of photonic bands through the interface states of a metasurface/photonic crystal, Phys. Rev. B 93, 041415 (2016).
  • (22) B. Midya and L. Feng, Topological multiband photonic superlattices, Phys. Rev. A 98, 043838 (2018).
  • (23) M. Maffei, A. Dauphin, F. Cardano, M. Lewenstein, and P. Massignan, Topological characterization of chiral models through their long time dynamics, New J. Phys. 20, 013023 (2018).
  • (24) S. Longhi, Probing topological phases in waveguide superlattices, Opt. Lett. 44, 2530 (2019).
  • (25) Z.-Q. Jiao, S. Longhi, X.-W. Wang, J. Gao, W.-H. Zhou, Y. Wang, Y.-X. Fu, L. Wang, R.-J. Ren, L.-F. Qiao, and X.-M. Jin, Experimentally Detecting Quantized Zak Phases without Chiral Symmetry in Photonic Lattices, Phys. Rev. Lett. 127, 147401 (2021).
  • (26) 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).
  • (27) K. Kawabata, K. Shiozaki, M. Ueda, and M. Sato, Symmetry and topology in non-Hermitian physics, Phys. Rev. X 9, 041015 (2019).
  • (28) E.J. Bergholtz, J.C. Budich, and F.K. Kunst, Exceptional topology of non-Hermitian systems, Rev. Mod. Phys. 93, 015005 (2021).
  • (29) K. Ding, C. Fang, and M. Guancong, Non-Hermitian topology and exceptional-point geometries, Nature Rev. Phys. 4, 745 (2022).
  • (30) Q. Wang and Y.D. Chong, Non-Hermitian Photonic Lattices: tutorial, arXiv:2212.00307v1 (2022).
  • (31) I. Rotter, A non-Hermitian Hamilton operator and the physics of open quantum systems, J. Phys. A 42, 153001 (2009).
  • (32) N. Moiseyev, Non-Hermitian quantum mechanics (Cambridge University Press, Cambridge, MA, 2011).
  • (33) Y. Ashida, Z. Gong, and M. Ueda, Non-Hermitian physics, Adv. Phys. 69, 3 (2020).
  • (34) C.M. Bender and S. Boettcher, Real spectra in non-Hermitian Hamiltonians having 𝒫𝒯\mathcal{PT} symmetry, Phys. Rev. Lett. 80, 5243 (1998).
  • (35) C.M. Bender, D.C. Brody, and H.F. Jones, Complex extension of quantum mechanics, Phys. Rev. Lett. 89, 270401 (2002).
  • (36) C.M. Bender, Making sense of non-Hermitian Hamiltonians, Rep. Prog. Phys. 70, 947 (2007).
  • (37) A. Mostafazadeh, Pseudo-Hermiticity versus 𝒫𝒯\mathcal{PT} symmetry: the necessary condition for the reality of the spectrum of a non-Hermitian Hamiltonian, J. Math. Phys. 43, 205 (2002).
  • (38) A. Mostafazadeh, Pseudo-Hermiticity versus 𝒫𝒯\mathcal{PT}-symmetry. II. A complete characterization of non-Hermitian Hamiltonians with a real spectrum, J. Math. Phys. 43, 2814 (2002).
  • (39) F. Cannata, G. Junker and J.Trost, Schrödinger operators with complex potential but real spectrum, Phys. Lett. A 246, 219 (1998).
  • (40) B. Bagchi, S. Mallik and C. Quesne, Generating complex potentials with real eigenvalues in supersymmetric quantum mechanics, Int. J. Mod. Phys. A 16, 2859 (2001).
  • (41) T.V. Fityo, A new class of non-Hermitian Hamiltonians with real spectra, J. Phys. A 35, 5893 (2002).
  • (42) O. Rosas-Ortiz, O Castanos, and D. Schuch, New supersymmetry-generated complex potentials with real spectra, J. Phys. A 48, 445302 (2015).
  • (43) J. Gong and Q. Wang, Stabilizing non-Hermitian systems by periodic driving, Phys. Rev. A 91, 042135 (2015).
  • (44) R. Yang, J.W. Tan, T. Tai, J.M. Koh, L. Li, S. Longhi, and C.H. Lee, Designing non-Hermitian real spectra through electrostatics, Sci. Bull. 67, 1865 (2022).
  • (45) C.M. Bender, 𝒫𝒯\mathcal{PT} symmetry in quantum physics: from a mathematical curiosity to optical experiments, EuroPhys. News 47, 17 (2016).
  • (46) R. El-Ganainy, K.G. Makris, M. Khajavikhan, Z.H. Musslimani, S. Rotter, and D.N. Christodoulides, Non-Hermitian physics and 𝒫𝒯\mathcal{PT} symmetry, Nat. Phys. 14, 11 (2018),
  • (47) L. Feng, R. El-Ganainy, and L. Ge, Non-Hermitian photonics based on parity-time symmetry, Nature Photon. 11, 752 (2017).
  • (48) S. Longhi, Parity-time symmetry meets photonics: A new twist in non-Hermitian optics, EPL 120, 64001 (2017).
  • (49) B. Midya, H. Zhao, and L. Feng, Non-Hermitian photonics promises exceptional topology of light, Nature Commun. 9, 2674 (2018).
  • (50) S.K. Ozdemir, S. Rotter, F. Nori, and L. Yang, Parity-time symmetry and exceptional points in photonics, Nature Mat. 18, 783 (2019).
  • (51) M.-A. Miri and A. Alù, Exceptional points in optics and photonics, Science 363, eaar7709 (2019).
  • (52) M.V. Berry, Physics of Nonhermitian Degeneracies, Czech. J. Phys. 54, 1039 (2004).
  • (53) W.D. Heiss, The physics of exceptional points, J. Phys. A 45, 444016 (2012).
  • (54) A. Mostafazadeh, Spectral singularities of complex scattering potentials and infinite reflection and transmission coefficients at real energies, Phys. Rev. Lett. 102, 220402 (2009).
  • (55) S Longhi, Spectral singularities and Bragg scattering in complex crystals, Phys. Rev. A 81, 022102 (2009).
  • (56) S. Longhi, Optical Realization of Relativistic Non-Hermitian Quantum Mechanics, Phys. Rev. Lett. 105, 013903 (2010).
  • (57) Ch. Miniatura, C. Sire, J. Baudon, and J. Bellissard, Geometrical Phase Factor for a Non-Hermitian Hamiltonian, EPL 13, 199 (1980).
  • (58) J. C. Garrison and E. M. Wright, Complex geometrical phases for dissipative systems, Phys. Lett. A 128, 177 (1988).
  • (59) G. Dattoli, R. Mignani, and A. Torre, Geometrical phase in the cyclic evolution of non-Hermitian systems, J. Phys. A 23, 5795 (1990).
  • (60) A. Mostafazadeh, A new class of adiabatic cyclic states and geometric phases for non-Hermitian Hamiltonians, Phys. Lett. A 264, 11 (1999).
  • (61) A.A. Mailybaev, O.N. Kirillov, and A.P. Seyranian, Geometric phase around exceptional points, Phys. Rev. A 72, 014104 (2005).
  • (62) H. Mehri-Dehnavi and A. Mostafazadeh, Geometric phase for non-Hermitian Hamiltonians and its holonomy interpretation, J. Math. Phys. 49, 082105 (2008).
  • (63) J. Gong and Q.-h. Wang, Geometric phase in 𝒫𝒯\mathcal{PT}-symmetric quantum mechanics, Phys. Rev. A 82, 012103 (2010).
  • (64) X.-D. Cui and Y. Zheng, Geometric phases in non-Hermitian quantum mechanics, Phys. Rev. A 86, 064104 (2012).
  • (65) J. Gong and Q.-h. Wang, Time-dependent 𝒫𝒯\mathcal{PT}-symmetric quantum mechanicsJ. Phys. A 46, 485302 (2013).
  • (66) S. Lin, X.Z. Zhang, and Z. Song, Amplitude control of a quantum state in a non-Hermitian Rice-Mele model driven by an external field, Phys. Rev. A 92, 012117 (2015).
  • (67) S. Lin and Z. Song, Non-Hermitian heat engine with all-quantum-adiabatic-process cycle, J. Phys. A 49, 475301 (2016).
  • (68) S. Longhi, Floquet exceptional points and chirality in non-Hermitian Hamiltonians, J. Phys. A 50, 505201 (2017).
  • (69) R. Hayward and F. Biancalana, Complex Berry phase dynamics in 𝒫𝒯\mathcal{PT}-symmetric coupled waveguides, Phys. Rev. A 98, 053833 (2018).
  • (70) J. Gong and Q.-h. Wang, Piecewise adiabatic following in non-Hermitian cycling, Phys. Rev. A 97, 052126 (2018).
  • (71) E.J. Pap, D. Boer, and H. Waalkens, A Unified View on Geometric Phases and Exceptional Points in Adiabatic Quantum Mechanics, SIGMA 18, 003 (2022).
  • (72) Y. Singhal, E. Martello, S. Agrawal, T. Ozawa, H. Price, and B. Gadway, Measuring the Adiabatic Non-Hermitian Berry Phase in Feedback-Coupled Oscillators arXiv:2205.02700 (2022).
  • (73) C. Dembowski, H.-D. Gräf, H. L. Harney, A. Heine, W. D. Heiss, H. Rehfeld, and A. Richter, Experimental Observation of the Topological Structure of Exceptional Points, Phys. Rev. Lett. 86, 787 (2001).
  • (74) S.-D. Liang and G.-Y. Huang, Topological invariance and global Berry phase in non-Hermitian systems, Phys. Rev. A 87, 012118 (2013).
  • (75) H. Schomerus, Topologically protected midgap states in complex photonic lattices, Opt. Lett. 38, 1912 (2013).
  • (76) H. Zhao, S. Longhi, and L. Feng, Robust Light State by Quantum Phase Transition in Non-Hermitian Optical Materials, Sci. Rep. 5, 17022 (2015).
  • (77) D. Leykam, K.Y. Bliokh, C. Huang, Y.D. Chong, and F. Nori, Edge modes, degeneracies, and topological numbers in non-Hermitian systems, Phys. Rev. Lett. 118, 040401 (2017)
  • (78) S. Lieu, Topological phases in the non-Hermitian Su-Schrieffer-Heeger model, Phys. Rev. B 97, 045106 (2018).
  • (79) R. Wang, X. Z. Zhang, and Z. Song, Dynamical topological invariant for the non-Hermitian Rice-Mele model, Phys. Rev. A 98, 042120 (2018).
  • (80) M. Pan, H. Zhao, P. Miao, S. Longhi, and L. Feng, Photonic zero mode in a non-Hermitian photonic lattice, Nat. Commun. 9, 1308 (2018).
  • (81) X.Z. Zhang and Z. Song, Partial topological Zak phase and dynamical confinement in a non-Hermitian bipartite system, Phys. Rev. A 99, 012113 (2019).
  • (82) F. Dangel, M. Wagner, H. Cartarius, J. Main, and G. Wunner, Topological invariants in dissipative extensions of the Su-Schrieffer-Heeger model, Phys. Rev. A 98, 013628 (2018).
  • (83) X. Ni, D. Smirnova, A. Poddubny, D. Leykam, Y. Chong, and A.B. Khanikaev, PT phase transitions of edge states at PT symmetric interfaces in non-Hermitian topological insulators, Phys. Rev. B 98, 165129 (2018).
  • (84) L.J. Lang, Y. Wang, H. Wang, and Y.D. Chong, Effects of non-Hermiticity on Su-Schrieffer-Heeger defect states, Phys. Rev. B 98, 094307 (2018).
  • (85) D.-J. Zhang, Q.-h. Wang, and J. Gong, Quantum geometric tensor in PT-symmetric quantum mechanics, Phys. Rev. A 99, 042104 (2019).
  • (86) C.-X. Du, N. Xu, L. Du, Y. Zhang, and J.-H. Wu, Topological edge states controlled by next-nearest-neighbor coupling and Peierls phase in a PT-symmetric trimerized lattice, Opt. Express 29, 37722 (2021).
  • (87) M. Ezawa, Non-Hermitian non-Abelian topological insulators with PT symmetry, Phys. Rev. Research 3, 043006 (2021).
  • (88) H. C. Wu, L. Jin, and Z. Song, Topology of an anti-parity-time symmetric non-Hermitian Su-Schrieffer-Heeger model, Phys. Rev. B 103, 235110 (2021).
  • (89) S. Tsubota, H. Yang, Y. Akagi, and H. Katsura, Symmetry-protected quantization of complex Berry phases in non-Hermitian many-body systems, Phys. Rev. B 105, L201113 (2022).
  • (90) S. Longhi, Bloch oscillations in complex crystals with 𝒫𝒯\mathcal{PT}-symmetry, Phys. Rev. Lett. 103, 123601 (2009).
  • (91) S. Longhi, Dynamic localization and transport in complex crystals, Phys. Rev. B 80, 235102 (2009).
  • (92) C. Elsen, K. Rapedius, D. Witthaut, and H.J. Korsch, Exceptional points in bichromatic Wannier-Stark systems, J. Phys. B 44, 225301 (2011).
  • (93) N. Bender, H. Li, F. M. Ellis, and T. Kottos, Wave-packet self-imaging and giant recombinations via stable Bloch-Zener oscillations in photonic lattices with local 𝒫𝒯\mathcal{PT} symmetry, Phys. Rev. A 92, 041803 (2015).
  • (94) M. Wimmer, M.-A. Miri, D. Christodoulides, and U. Peschel, Observation of Bloch oscillations in complex 𝒫𝒯\mathcal{PT}-symmetric photonic lattices, Sci. Rep. 5, 17760 (2015).
  • (95) Y.-L. Xu, W.S. Fegadolli, L. Gan, M.-H. Lu, X.-P. Liu, Z.-Y. Li, A. Scherer, and Y.-F. Chen, Experimental realization of Bloch oscillations in a parity-time synthetic silicon photonic lattice, Nat. Commun. 7, 11319 (2016).
  • (96) S. Longhi, Non-Bloch-Band Collapse and Chiral Zener Tunneling, Phys. Rev. Lett. 124, 066602 (2020).
  • (97) C. Qin, B. Wang, Z.J. Wong, S. Longhi, and P. Lu, Discrete diffraction and Bloch oscillations in non-Hermitian frequency lattices induced by complex photonic gauge fields, Phys. Rev. B 101, 064303 (2020).
  • (98) S. Xia, C. Danieli, Y. Zhang, X. Zhao, H. Lu, L. Tang, D. Li, D. Song,and Z. Chen, Higher-order exceptional point and Landau-Zener Bloch oscillations in driven non-Hermitian photonic Lieb lattices, APL Photonics 6, 126106 (2021).
  • (99) L. Zheng, B. Wang, C. Qin, L. Zhao, S. Chen, W. Liu, and P. Lu, Chiral Zener tunneling in non-Hermitian frequency lattices, Opt. Lett. 47, 4644 (2022).
  • (100) S. Longhi, Non-Hermitian Bloch-Zener phase transitions, Opt. Lett. 47, 6345 (2022).
  • (101) S. Yao and Z. Wang, Edge states and topological invariants of non-Hermitian systems, Phys. Rev. Lett. 121, 086803 (2018).
  • (102) B. M. Breid, D. Witthaut, and H.J. Korsch, Bloch-Zener oscillations, New J.Phys. 8, 110 (2006).
  • (103) S. Longhi, Optical Zener-Bloch oscillations in binary waveguide arrays, EuroPhys. Lett. 76, 416 (2006).
  • (104) S. Longhi, Bloch oscillations and Zener tunneling with nonclassical light, Phys. Rev. Lett. 101, 193902, (2008).
  • (105) F. Dreisow, A. Szameit, M. Heinrich, T. Pertsch, S. Nolte, A. Tünnermann, and S. Longhi, Bloch-Zener Oscillations in Binary Superlattices, Phys. Rev. Lett. 102, 076802 (2009).
  • (106) S. Kling, T. Salger, C. Grossert, and M. Weitz, Atomic Bloch-Zener Oscillations and Stückelberg Interferometry in Optical Lattices, Phys. Rev. Lett. 105, 215301 (2010).
  • (107) S. Longhi, Bloch-Zener quantum walk, J. Phys. B 45, 225504 (2012).
  • (108) Y. Zhang, D. Zhang, Z. Zhang, C. Li, Y. Zhang, F. Li, Mi. R. Belic, and M. Xiao, Optical Bloch oscillation and Zener tunneling in an atomic system, Optica 4, 571 (2017).
  • (109) Y. Sun, D. Leykam, S. Nenni, D. Song, H. Chen, Y.D. Chong, and Z. Chen, Observation of Valley Landau-Zener-Bloch Oscillations and Pseudospin Imbalance in Photonic Graphene, Phys. Rev. Lett. 121, 033904 (2018).
  • (110) S. Longhi, Topological phase transition in non-Hermitian quasicrystals, Phys. Rev. Lett. 122, 237601 (2019).
  • (111) S. Longhi, Metal-insulator phase transition in a non-Hermitian Aubry-Andre-Harper model, Phys. Rev. B 100, 125157 (2019).
  • (112) S. Weidemann, M. Kremer, S. Longhi, and A. Szameit, Topological triple phase transition in non-Hermitian Floquet quasicrystals, Nature 601, 354 (2022).
  • (113) R. Uzdin, A. Mailybaev, and N. Moiseyev, On the observability and asymmetry of adiabatic state flips generated by exceptional points, J. Phys. A 44, 435302 (2011).
  • (114) M.V. Berry and R. Uzdin, Slow non-Hermitian cycling: exact solutions and the Stokes phenomenon, J. Phys. A 44, 435303 (2011).
  • (115) H. Wang, L.J. Lang, and Y.D. Chong, Non-Hermitian dynamics of slowly varying Hamiltonians, Phys. Rev. A 98, 012119 (2018).