In the present paper, we aim to firmly establish the adiabatic properties of two-level non-Hermitian quantum structures evolving along generalized (open/acyclic or closed/cyclic) paths in parameter space. Analytical solutions in terms of Airy and modified Bessel functions have been retrieved for linear and hyperbolic temporal dependencies in parity-time-symmetric-like systems, which were subsequently studied in the slowly varying limit to show conversion to one of the instantaneous eigenstates. Such a mode switching behavior is found to be an identifying feature of dissipative quantum settings, whether they evolve along cyclic or acyclic trajectories, and this has been proven in our paper by separately analyzing the dynamics of (i) the ratio of the state vector components, via a variant of the Möbius transformation, and (ii) the complex probability amplitudes, through a systematic inspection of the mode population equations. In the latter instance, it was furthermore shown that the identity of the eigenstate, to which the quantum arrangement transitions, depends highly on the magnitude of the adiabatic rate β. Along these lines, the concepts of the instantaneous and averagely dominant eigenstates are brought forth, while a reconfigurable photonic switch is also proposed, which can convert either to the or to the modes based on the total period of evolution. Finally, we apply our findings in the case of closed parametric paths to demystify the recently reported symmetric and asymmetric state conversion effects and additionally demonstrate that operation at or near exceptional points does not qualitatively affect the conclusions of the current investigation.
I. INTRODUCTION
Non-Hermitian quantum mechanics constitutes an important extension to the standard Hermitian Hamiltonian formalism, as it enables us to investigate the dynamics of systems (S) that interact with the environment (E) and thus exhibit quantum dissipation. In this case, we say that the arrangement under study (S) suffers from decoherence (i.e., a definite phase relation no longer exists between the different quantum states), which in turn leads to information loss (from S to E) and to the generation of entanglement between the system and its surroundings. To overcome this issue and obtain a problem that can be still described by Hermitian Hamiltonians, we could have simply considered S and E together as a single unit. Following this thought process incrementally, it would eventually require having knowledge of the state of the entire universe, which even desirable is not possible. For practical purposes, instead of focusing on an infinite number of degrees of freedom, it is preferable to limit our attention to a smaller number of relevant variables describing the subsystem S and subsequently attain the corresponding effective equations of motion. The latter take into account any irrelevant degrees of freedom (associated with the interaction between S and E) via appropriate stochastic or dissipative terms. It should also be noted that decoherence can arise not only due to energy exchange with the environment but as a direct consequence of the act of measuring S (quantum decoherence/measurement theory/dissipation/entanglement are all highly interrelated concepts1,2). Asides from all the previous discussion on the importance of open quantum arrangements, it is noteworthy to mention that certain quantum dynamical problems (e.g., decay or capture resonance effects and electron/photon scattering from atoms or molecules) can be more efficiently and comprehensively treated when employing outgoing/incoming wave boundary conditions, which by itself demands the definition of non-Hermitian Hamiltonians3 (their Hermitian counterparts require that the wavefunction vanishes at the boundary/infinity). Of course, the applicability of non-Hermitian theory is by no means limited to the quantum domain but rather extended to a plethora of other fields dealing with non-conservative or stochastic settings, including biological physics, photonics, optomechanics, hydrodynamics, and neural networks.3,4
In recent years, a flurry activity has been noticed in the field of non-Hermitian physics and this has been greatly incited from the seminal paper by Bender and Boettcher5 on parity-time -symmetric Hamiltonians (i.e., Hamiltonians H that commute with the parity-time operator according to ). Such non-conservative settings exhibit a spontaneous symmetry breaking point, below and above which the eigenspectra become entirely real (the symmetry of Hamiltonian H is then said to be unbroken, given that guarantees that and H share a simultaneous set of eigenstates) and entirely complex (the symmetry of H is then broken, given that no longer implies that and H share the same eigenstates owing to the antilinearity of —parity operator : linear; time reversal operator : antilinear), respectively. An even greater peculiarity is that at this symmetry breaking threshold, eigenvectors and eigenvalues coalesce; such a degeneracy, which is a characteristic attribute of non-Hermitian structures, is known in the literature as an exceptional point (EP) and has led to a number of intriguing observations.6–9 Latest research has focused on the slowly varying properties of these open Hamiltonian arrangements. According to the standard quantum adiabatic theorem, as originally developed by Born and Fock10–12 in the context of Hermitian theory, a physical setting is expected to remain in its instantaneous eigenstate assuming that any external influences act on it slowly enough. Yet the previous statement is violated within non-conservative environments, where it has been shown that mode switching occurs for evolution along cyclic paths in the Hamiltonian parameter space13–22 and this has been exploited for the demonstration of asymmetric transport in a variety of classical and quantum experimental setups.23–28 Here, we extend these findings for acyclic parametric trajectories and more importantly unveil a previously overlooked fact: the identity of the mode describing the resulting quantum state highly depends on the magnitude of the adiabatic rate of evolution β. Capitalizing on the latter, we shall elaborate on the results reported in cyclic non-Hermitian configurations and also clear out any prior misconceptions regarding the correlation between EP encirclement and symmetric (SSC)/asymmetric (ASC) effects—more specifically, it will be seen that encircling EPs is neither a necessary nor a sufficient condition for SSC or ASC to take place.
Overall, the present article establishes the fundamental principles of open quantum systems, pertaining to their adiabatic evolution behavior (in what follows, the terms adiabatic, slowly varying, and quasistatic shall be used equivalently and interchangeably). While here we emphasize on two-level Hamiltonians, the physical insights and mathematical reasoning can be extended to higher-dimensional Hilbert spaces (particular examples related to 3 × 3 non-Hermitian Hamiltonians are included in the supplementary material; however, this topic will be explored in greater depth in our future research). Along these lines, the dynamical equations dictating the quantum state vector evolution associated with a 2 × 2 generalized and time-dependent Hamiltonian are initially retrieved (Sec. II A). Appropriate state vector transformations are also discussed, with the purpose of broadening the range of analytically solvable scenarios. Via the transfer matrix method, it is shown how the full system dynamics can be attained at any time instant, and this is subsequently applied in the special case of -symmetric-like Hamiltonian arrangements (Sec. II B). The parameter space is then defined by the gain/loss and detuning factors, and in this respect, two different types of parametric paths were examined: linear (this leads to parabolic cylinder and Airy solutions, with the latter applicable under certain approximations—Sec. II B 1) and hyperbolic (the dynamics can then be expressed in terms of modified Bessel functions—Sec. II B 2). The slowly varying response was studied in both circumstances, by employing the asymptotic properties of the corresponding solutions. Under an acyclic variation of the system parameters, it was illustrated that the resulting quantum state is always described by one of the instantaneous eigenstates of the non-Hermitian Hamiltonian, regardless of any initial excitation conditions. This phenomenon, known as state conversion or mode switching, is intimately linked to the Stokes phenomenon of asymptotics, as will also become evident through our detailed mathematical calculations.
In Secs. III and IV, we are concerned with the quasistatic features of generalized non-Hermitian settings [the respective trajectories within the parameter space can assume arbitrary geometric forms and may be either open (Sec. III) or closed (Sec. IV)]. In this regard, our study considers separately the evolution of (i) the ratio of the state vector components (Sec. III) and (ii) the mode population amplitudes (Sec. III and Appendix B 3). A discretization of the Schrödinger equation into small time steps indicates that quantity (i) obeys a time-dependent version of the Möbius transformation. By leveraging on the properties of the latter (after an appropriate adaptation for the time-dependent case), it is shown that under non-Hermitian and sufficiently slowly varying conditions, the quantum state converts to the least dissipative eigenstate except for very special circumstances (in the Hermitian case, the standard quantum adiabatic theorem is confirmed). Our derivation generalizes the findings of our recent work,22 where it is theoretically demonstrated that for small time steps, state conversion occurs in discrete non-Hermitian structures for a cyclic variation of the system parameters.
Regarding now quantity (ii), the quantum state is expressed at each time instant in terms of the instantaneous eigenmodes, in a manner also consistent with Wentzel–Kramers–Brillouin (WKB) theory for very large evolution periods (the WKB theory results were additionally utilized to confirm the asymptotic behavior of the analytical solutions retrieved in Secs. II B 1 and II B 2, in the context of -symmetric-like configurations). A rigorous investigation of the mode population equations then showed that the resulting system dynamics will be governed by a single eigenstate, whose identity depends on the degree of adiabaticity (the role of the non-adiabatic mode couplings has also been clearly highlighted). In this respect, the concepts of instantaneous/averagely dominant eigenstates were employed to describe moderately/extremely slow non-Hermitian processes with a special emphasis on the topology of the associated eigenspectra evolution graphs. This observation was subsequently used to propose a reconfigurable optical switch based on discrete photonic lattices, which can operate in either a cyclic or an acyclic fashion—in such a classical optics platform, the state is defined through the polarization of the propagating field, its dynamics are dictated by a discrete Schrödinger-like equation, while the adiabatic rate of evolution is controlled via the number of round trips along an appropriately designed fiber loop (for more technical details on the Schrödinger-type equations governing discrete or continuous photonic setups, see Refs. 22, 28, and 29). The case of closed paths in parameter space was examined in a separate section (Sec. IV) in order to (a) provide a more detailed interpretation of the symmetric/asymmetric mode switching effects observed within non-conservative environments13–28 (here, in essence, we apply the analytic findings of Sec. III pertaining to parametric trajectories of arbitrary geometric form) and (b) demonstrate the functional characteristics of the aforementioned optical switch, showcasing its capacity to support a dual operational regime (SSC or ASC) determined by the value of β. Finally, it has been shown that the crossing of spectrum degeneracies can have only a quantitative (magnitude of β for conversion to take place) rather than a qualitative impact in our study.
II. ASYMPTOTIC ANALYSIS OF EXACTLY SOLVABLE OPEN QUANTUM SYSTEMS
A. Two-level Hamiltonian settings
Conventional Hermitian arrangements satisfy the condition H†(t) = H(t) or equivalently M†(t) = −M(t) [Q†(t) refers to the conjugate transpose/adjoint of Q(t)], which in turn demand that the eigenvalues/eigenenergies of the Hamiltonian matrix H(t) are purely real and that the eigenvalues λ±(t) of the system matrix M(t) are purely imaginary [symbols (+) and (−) designate the modes of the two-level Hamiltonian structure under study]. Here, we are particularly interested in the non-Hermitian case (open quantum systems), where the eigenvalues are in general complex. The imaginary (real) part of [λ±(t)] may signify (i) gain/loss in optical setups and (ii) dissipation/decoherence in quantum mechanical setups. Existing analytical or experimental investigations into state conversion effects in slowly varying non-conservative settings13–28 predominantly rely on a cyclic variation of the system parameters. In the current paper, such a condition is relaxed, i.e., we allow the Hamiltonian arrangement to move along open paths in parameter space .
B. Traceless and symmetric non-Hermitian Hamiltonian model
It is interesting now to note that Eq. (15b) corresponds to a harmonic oscillator exhibiting a complex and time-dependent damping factor [2f(t)] with a positive or a negative real profile, depending on the sign of g(t) = Re[f(t)] (Re[·] and Im[·] will designate the real and imaginary parts of the associated complex quantities, respectively). This observation is essential, since it enables us to take advantage of the large amount of analytical studies (and known analytical solutions) for damped harmonic oscillator systems, by simply identifying the right temporal dependencies for f(t). In this regard, if we allow f(t) to vary linearly [f(t) = ηf,1t + ηf,0 with ηf,1, ηf,0 constants] or hyperbolically [] with time, then Eq. (15b) can be immediately recognized to be of the Hermite- and Bessel-type, respectively [in Ref. 22, an even richer landscape of dynamics is attained for alternative temporal dependencies of f(t)]. In what follows, both of the aforementioned scenarios shall be considered in the asymptotic regime—more specifically, Eq. (15a) shall be used when f(t) = ηf,1t + ηf,0 (this will lead to the less familiar parabolic cylinder functions and under certain assumptions to the Airy functions) and Eq. (15b) shall be employed when .
So far, our discussion has been purely theoretical, but it should be highlighted that time-dependent non-Hermitian models of the form shown in Eq. (14) have been realized in a variety of experimental settings.23–28 What is even more noteworthy is the demonstrated capability of precisely tailoring the complex damping factor via electro-optic amplitude (tuning of the gain coefficient) and phase (tuning of the phase mismatch) modulator components within judiciously designed fiber loop networks28 (in this case, the state of the system is described by the polarization of the propagating field). This in turn opens up the possibility of emulating a large range of physical arrangements obeying dynamics analogous to Eqs. (14)–(17), including the structures that will be analyzed in Secs. II B 1 and II B 2.
1. Linear time dependence for complex damping factor f(t)
2. Hyperbolic time dependence for complex damping factor f(t)
III. ADIABATIC EVOLUTION PROPERTIES OF GENERALIZED HAMILTONIAN SETTINGS
So far, we have demonstrated the potential of Eq. (49) in describing the limiting behavior of differential equations of a form analogous to Eq. (3), in alignment also with the WKB method analyzed in Appendix A 2 (the formulas and analysis can be directly extended to higher-order differential equations, ascribed to higher-dimensional Hamiltonian configurations). Of interest is now to investigate the circumstances where one of the two terms appearing in the right-hand side of Eq. (49) always becomes prevalent, independently of the initial excitation conditions. This physically implies that the quantum response will be dictated by one of the two instantaneous eigenmodes [either or ], and therefore, state conversion will have occurred. Evidently, the aforementioned scenario cannot emerge in Hermitian structures, since will be purely imaginary and thus both terms in Eq. (49) will be purely oscillatory and of a similar magnitude (see Appendix B 1). This is also in agreement with conventional quantum adiabatic theory, according to which if an arrangement is excited at one of its instantaneous eigenstates, then it remains in such an eigenstate as long as any external influences act on it slowly enough (the latter equivalently suggests that the response will depend on the initial conditions).
It would be of particular interest at this moment, to physically interpret our choice to correlate (i.e., the “fixed points” of ) with eigenvectors . In this vein, we can consider the relation in conjunction with the conventional adiabatic theorem underpinning Hermitian quantum processes: if the instantaneous eigenstates are originally excited , then these will continue to dictate the state vector dynamics in the slowly varying limit . In the context of non-Hermitian settings, the latter statement will not uphold, since are not any more indifferent (i.e., they become attractive/repulsive—see the subsequent analysis), and therefore, small variations in the arguments of the -transform can cumulatively lead to large deviations in the output.
Under Hermitian conditions , are neither attractive nor repulsive but indifferent and the -transform can be classified as elliptic.
Under non-Hermitian driving, if for all possible values of n, then the conclusions in (i) are also relevant here.
Under non-Hermitian driving [which in general implies ], if , then is repulsive (attractive) and the transform is said to be loxodromic [if , the roles of are reversed—in Appendix E, it is analytically shown how the attractive/repulsive behavior of emerges in loxodromic transformations].
From the preceding observations, we can infer that any initially excited quantum state will convert to the instantaneous dominant eigenmode [, if ] during adiabatic evolution [or mathematically stated, the iterative action of the -transform will result in , with symbol ◦ indicating function composition and r0 given by initial conditions], as long as non-Hermiticity and an imbalance between the real parts of are both present. This in turn confirms the detailed investigation of Appendix B 3, at least when superadiabatic rates β are considered. Moreover, the impact of the various types of -transform on the evolution traits of rn has been numerically demonstrated in Fig. 3 and is in agreement with our theoretical predictions. Finally, the reader can refer to the supplementary material for a generalization of our approach to three-level Hamiltonian systems, whereby the properties of the two-dimensional -transform have been employed.
IV. SYMMETRIC AND ASYMMETRIC STATE CONVERSION UNDER THE ACTION OF CYCLIC AND SLOWLY VARYING NON-HERMITIAN HAMILTONIANS
Having provided an extensive qualitative and quantitative insight as to the quasistatic response of two-mode non-Hermitian quantum configurations traversing trajectories of any shape in space, we are now in the position to apply our findings in the case of closed paths . Before doing so, it is necessary to first specify the meaning of symmetric (SSC) and asymmetric (ASC) state conversion in the context of cyclic Hamiltonian arrangements: in the former/latter instance, clockwise (CW) and counterclockwise (CCW) encirclements within space lead to switching to the same/different instantaneous eigenstates after a full cycle, independently of how the system was initially prepared. Along these lines, and leveraging on the analysis of Sec. III and Appendix B 3, it can be deduced that for superadiabatic parameter changes, which entail conversion to mode , SSC will always be observed if (T: cyclic evolution period)—this becomes self-evident, since the eigenstate exhibiting a larger value for Re[λ(T)] (and therefore recognized as the mode) is uniquely determined and will not depend on the encircling direction. The only situation where ASC might possibly arise is when , but then cannot be strictly defined at t = T. This can be resolved by simply inspecting, just before time T, which of the eigenmode branches is characterized by instantaneous eigenvalues with a larger real part—such a branch will eventually dominate for sufficiently large T. The aforementioned rationale mathematically implies that the dynamical behavior of the quantum system is examined at the limit t → T, which in turn based on the theory of limits and the continuity of the underlying functions yields an accurate description of what happens at t = T. {For a graphical representation of the latter, we direct the reader to Figs. 6(a) and 6(b), where the dynamics along SE [ at E] are explored in terms of path SK, as point K approaches E.} A completely analogous reasoning can also be employed in the presence of eigenspectrum degeneracies, whether they occur at the end or anywhere along [see Figs. 6(c)–6(h)].
To substantiate our discussion thus far, separate numerical calculations have been performed, and the results are shown in Figs. 4–6 for the non-Hermitian model of Sec. II B, assuming rhombic parametric trajectories . In this vein, ASC was noticed in the first and third rows of Fig. 4 [see Figs. 4(b) and 4(f)], where the ending point of is located to the left of the EP degeneracy at (κ, 0), thereby leading to . {The -symmetric-like system under study exhibits eigenvalues . Therefore, the relation can be satisfied iff .} Meanwhile, in the second row of Fig. 4 and for all the simulation scenarios in Fig. 5, SSC was attained as then . (Here, we refer to the main panels of the last column of Fig. 5, where conversion to the mode takes place.) It should be highlighted that in all the aforementioned examples, switching to mode is observed for both CW and CCW parametric encirclements. This aligns with the analytical investigations of Sec. II B, where the quantum state was found to be dictated by the mode branch in the extreme superadiabatic limit T → ∞ [the findings of Sec. II B have been numerically verified through Figs. 1, 3(a)–3(c), 6(a), and 6(b), bearing also in mind that in these instances, as detailed in the caption of Fig. 6]. Such a dynamical behavior persists, even when EPs are crossed by —this is illustrated in Figs. 6(c)–6(h) [see the inset of Fig. 6(f)], with the state vector dynamics along being examined in terms of a family of auxiliary trajectories , which do not pass through any spectrum degeneracies (see the caption of Fig. 6, for a more comprehensive description).
We shall turn now our attention to the simulation results depicted in the third column of Fig. 5, where conversion to the averagely dominant mode occurs. [We need not further analyze the graphs of Fig. 4 with respect to , as either superadiabaticity is reached and thus it cannot emerge (second and third rows of Fig. 4), or simply (first row of Fig. 4) and therefore does not alter in any way the resulting dynamics.] Relative to the fourth column, period T has been reduced, so as the topology of the entire path has an effect on the identity of the dominating eigenstate. Consequently, instead of getting a SSC-type response for all parameter configurations, asymmetric switching is noticed in the first and second rows of Fig. 5. Notably, conversion to the mode can still manifest in the presence of EPs, as evidenced in Figs. 6(c)–6(f). {Figures 6(e) and 6(f) [see the main graph of Fig. 6(f)] provide a more illustrative example of the latter, since then .}
Bringing it altogether, an important and previously overlooked dynamic trait of slow non-Hermitian cycling is unveiled: the capacity to promote reconfigurable ASC/SSC operation by tuning the overall period of evolution. For instance, if T is increased for the scenarios demonstrated in the first two rows of Fig. 5, the switching response shifts from ASC (third column) to SSC (fourth column). An even more intricate switching pattern is attained in the last row of the same figure, as T grows larger: SSC [Fig. 5(k)] → ASC [the inset of Fig. 5(l)] → SSC [the main panel of Fig. 5(l)]. To interpret these peculiar (yet extremely useful) dynamical features, we need to consider that traversing a path in the forward (CW) and backward (CCW) manner will produce different state vector evolutions [this can be directly seen via Eq. (54), where in general ΞNΞN−1⋯Ξ0 ≠ Ξ0⋯ΞN−1ΞN given the non-commutativity of matrix multiplication], except from very special cases [e.g., if H(t) = H(T − t), M(t) = M(T − t) in Eqs. (1) and (2), or equivalently if Ξn = ΞN−n, Hn = HN−n, Mn = MN−n in Eq. (54), or when ASC occurs in cyclic non-Hermitian settings]. In this regard, and assuming that conditions for the emergence of are met, will be characterized by two distinct critical values, and , signifying the thresholds below/above which the quantum state converts to , with respect to the two opposite traversal directions. The preceding suggests that a maximum of three different (and neighboring) ranges of T can arise, each resulting in ASC or SSC responses. The exact number of such intervals varies according to whether or under CW/CCW conditions, and if mode is supported or not. Along these lines, a variety of switching patterns can be acquired, depending on the chosen topology of .
The foregoing arguments ultimately lead to the proposition for a reconfigurable optical switch/omnipolarizer based on fiber loop networks. Such synthetic photonic lattices, which were outlined in Sec. II B, allow for the realization of the continuous two-level Schrödinger equation depicted in Eq. (1) or (2), but in a discrete time fashion [see Eq. (54)]. (In what follows, we shall briefly describe the operation of fiber loop setups, but for more technical details, the reader should resort to the supplementary material of Ref. 22 or Ref. 28.) The state vector will correspond to the polarization components Ex, Ey of the propagating field, which can be coupled via a polarization controller (implemented through a variable retardation wave-plate or a birefringent fiber) and individually manipulated by intensity (amplifying/attenuating Ex, Ey) and phase modulators (imparting phase on Ex, Ey). Beam splitters separate the polarization components for independent modulation, while beam combiners merge them back together. Each round trip along the fiber loop marks an incremental time step (tn → tn+1, tn+1 − tn = Δ), and thus, the adiabatic rate will be dictated by the overall number of round trips N (β = 1/T ∝ 1/N, where T = NΔ and Δ is a fixed small time step). The great degree of tunability inherent in these discrete optical configurations enables the implementation of diverse topologies, thereby giving rise to a wide range of switching patterns [the exact type of response (ASC or SSC) can be specified, after appropriately adjusting N]. While this paragraph focuses on a particular class of photonic networks, our findings have broader applicability to quantum or other physical settings as long as experimental platforms with comparable levels of modulation flexibility are available.
To summarize, we would like to make the following remarks:
The adiabatic transport features underpinning the non-Hermitian evolution can be determined by examining the topology of parametric contour , either at (or near) the end, or in an average fashion. In the former (latter) occasion, which arises for sufficiently large (moderate) values of T, we need to compare quantities to identify the prevalent eigenstate. The above criteria pertain to both open and closed trajectories in space and are consistent with our recent study in Ref. 22, where the state vector dynamics of discrete non-Hermitian cyclic arrangements were analyzed but only in the superadiabatic limit.
Exceptional point encirclement is neither a necessary [see Figs. 4(e) and 4(f)] nor a sufficient [see Figs. 4(c) and 4(d)] condition for asymmetric state conversion to appear. This holds true for symmetric switching as well [compare Figs. 4(a) and 4(b) with Figs. 4(c) and 4(d)]. In Appendix C, a separate investigation has been conducted as to the effect of EP encirclement on the resulting switching performance.
The conclusions of the current analysis will continue to apply in the presence of points where eigenstates or are not well-defined, as inferred from Fig. 6. The identities of and can then be specified, by inspecting quantities and in the limit t → T. Furthermore, the crossing of EPs is not prohibitive for any state conversion effects to take place (of course, operation at such degeneracies is highly unlikely in actual experimental situations, owing to their inherent sensitivity34–36). All the preceding circumstances are reviewed in a more detailed manner within Appendix B 3, in terms of the mode population equations.
The technique, presented in Fig. 6, can be utilized to investigate/verify the dynamics along arbitrary parametric paths, at least when non-Hermitian conditions are considered and mode switching is noticed (this in turn signifies the robustness of non-Hermitian switching processes). Yet, if mode switching does not occur, the output state will typically vary significantly even with small changes in [the previous can be deduced, for instance, from Figs. 3(d) and 3(e), where the state vector fluctuates considerably with time], and then, the methodology of Fig. 6 may not offer distinct advantages.
In the current section, we explored the potential of attaining reconfigurable ASC/SSC operation within cyclic non-Hermitian environments, by exploiting the adiabatic state transition to modes or contingent upon the magnitude of T (this subsequently led to our proposal for reconfigurable omnipolarizer devices). An analogous functionality is supported in acyclic non-conservative settings, but then, it is not pertinent to discuss about (a)symmetric state conversion, as the starting and ending points in space do not overlap. Essentially, in the case of open parametric trajectories, we examine the state vector response as they are traversed in the forward and backward directions, which can also yield a rich variety of switching motifs.
The -symmetric-like Hamiltonian, introduced in Sec. II B and used in the numerical (and some of the theoretical) computations, serves as a prototypical model upon which much of the existing literature on non-Hermitian structures is based. However, the implications of the present study extend beyond such a framework, to non-Hermitian Hamiltonians of arbitrary functional forms, as evidenced from our systematic analytical derivations. Furthermore, the reported switching to the or modes is independent of the initial excitation (here, we assumed random initial conditions, yet the results remain unaffected if the instantaneous eigenstates get excited) and the precise shape of the traversed path in space [in the various numerical scenarios, we emphasized on linear parametric trajectories and linear time dependencies, but similar observations hold for alternative path profiles or temporal evolutions (see Fig. 1)].
V. CONCLUSIONS AND OUTLOOK
In this paper, the dynamics of time-dependent two-level open quantum systems, both cyclic and acyclic, have been rigorously investigated in the slowly varying limit for parametric trajectories of arbitrary geometries. Such an analysis has been conducted both in a discrete (state vector component ratio) and in a continuous time fashion (mode population equations), and it has been shown that conversion to the least dissipative instantaneous eigenstate is expected for superadiabatic processes (special emphasis is then given to the topology of the respective eigenmode branches toward the end of the parametric evolution). This has been confirmed in the particular case of traceless and symmetric non-Hermitian (or -symmetric-like) Hamiltonian configurations, where exact solutions have been retrieved for two distinct types of time dependencies (linear/hyperbolic—Airy/modified Bessel solutions). Our asymptotic calculations for very large evolution periods have clearly demonstrated the role of the Stokes phenomenon in the observed mode switching behavior. While a two-dimensional space was assumed in the aforementioned scenario (defined by the gain and detuning coefficients), the main analysis and conclusions of our work (Secs. II A, III, and IV) are not constrained by the dimensionality of the parameter space.
An important outcome of the current study is that there exists a transitional range of values for the adiabatic rate β, below and above which conversion still takes place but not necessarily to the same eigenstate. This led to the introduction of the averagely dominant eigenstate , whose identity is determined after examining the topology of the eigenmode branches as a whole and not only toward their ending points (the competition for dominance between and has been illustrated within Appendix B 3 via a detailed inspection of the mode population equations, while also highlighting the significance of the non-adiabatic mode couplings). All our findings have been eventually applied in the case of closed trajectories in space, to provide a reasoning behind the recently reported SSC and ASC effects in cyclic non-conservative Hamiltonian arrangements. Moreover, we have proposed a reconfigurable optical switch/omnipolarizer, which under cyclic (acyclic) conditions can exhibit either SSC or ASC behaviors (can switch either to the or to the modes) by tuning the overall period of evolution.
To conclude this discussion, even though the great majority of the aspects concerning the adiabatic properties of two-mode non-Hermitian Hamiltonian settings have been elucidated, there are still certain topics that need to be separately addressed. One of them is to derive more accurate estimates for the critical threshold of β that signifies (i) the transition of the quantum state response from to (βcr) and (ii) the onset of any state conversion phenomena (βad). Along these lines, a question arises as to whether it is possible under certain circumstances for βad < βcr, which consequently implies that cannot emerge (the quantum state can convert then only to the eigenstate). Furthermore, it should be noted that while the physical intuition and results presented here can be readily extended to higher (and even infinite)-dimensional Hilbert spaces, the underlying mathematical techniques should be appropriately adapted in such situations. Actually, this will be the subject of forthcoming investigations; already though, some characteristic examples involving three-level non-Hermitian Hamiltonians have been provided in the supplementary material. Finally, of great practical interest is to explore the impact of any nonlinear (or other external perturbation) effects on the resulting adiabatic transport dynamics, as this will enable a more systematic evaluation of the performance and robustness of mode switching platforms inspired from non-Hermitian quantum theory.
SUPPLEMENTARY MATERIAL
The supplementary material provides details as to how the developed methodologies can be extended to higher-dimensional Hamiltonian settings. In this respect, the specific case of 3 × 3 non-Hermitian Hamiltonians has been analytically examined in the superadiabatic limit, in terms of both the mode population and the state vector component ratio dynamics (in the latter instance, a two-dimensional generalization of the G-transform has been utilized). Switching to the instantaneous dominant eigenmode D has been shown then, under both cyclic and acyclic conditions, given an imbalance among the real parts of the system eigenvalues λ(t).
ACKNOWLEDGMENTS
N.S.N. was supported by the Bodossaki Foundation.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Nicholas S. Nye: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Software (lead); Validation (equal); Writing – original draft (lead); Writing – review & editing (equal). Nikolaos V. Kantartzis: Validation (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.
APPENDIX A: SLOWLY VARYING BEHAVIOR OF GENERALIZED QUANTUM HAMILTONIAN SETTINGS IN TERMS OF THE WKB METHOD
1. Calculating the eigenvalues and the left and right eigenvectors of a two-level Hamiltonian system
We shall conclude the present section by making the following comments:
Here and overall in our analysis, we emphasize on the system matrix M(t) instead of the Hamiltonian matrix H(t), owing to the non-Hermiticity of the quantum arrangements under study [H†(t) ≠ H(t) or equivalently M†(t) ≠ −M(t), given that M(t) = −iH(t)]. This allows us to turn our attention to the real part of the eigenvalues λ±(t) of M(t), instead of the imaginary part of the eigenvalues of H(t) (known in the literature as eigenenergies), and thus avoid the multiple appearances of the imaginary unit i, which can be rather confusing. Of course, since M(t) and H(t) are linearly related [M(t) = −iH(t)], the same will be true for the associated eigenvalues with the left and right eigenvectors remaining unchanged. (Similarly, for the discrete evolution studied in Sec. III and further analyzed in Appendixes D and E, it applies that Ξn = I + MnΔ = I − iHnΔ, which in turn implies that matrices Ξn, Mn, Hn share the same eigenvectors and for the respective eigenvalues.)
- In Hermitian settings [H†(t) = H(t) or M†(t) = −M(t)], the eigenvalues λ±(t) become imaginary [eigenenergies are real], while the left and right eigenvectors are conjugate transposes of each other,{In the above formula, stands for the complex conjugate of λ±(t).} Moreover, for spectrum degeneracies or diabolic points to arise , it comes from Eq. (A10) that(A18)where we have applied relations , with and (, while index DP signifies operation at the diabolic point). Equation (A19) can hold iff p = q and or equivalently MDP = ipI (and HDP = −pI for the Hamiltonian matrix). In this case, the system eigenvalues coincide at λDP = ip, but a fundamental difference can be noticed with respect to non-Hermitian degeneracies: whereas at EPs, a coalescence of eigenvectors takes place, this does not occur at DPs. Stated in terms of vector spaces, the dimensionality of the eigenspace (i.e., the vector space formed by the eigenvectors associated with a particular eigenvalue) at EPs is one, while at DPs is two, at least for the two-mode systems that we examine. [The only circumstance in which the dimensionality of the eigenspace at a non-Hermitian degeneracy is two (deeming it thus as a DP with non-coalescent eigenvectors) is when the system matrix M/Hamiltonian H becomes proportional to the identity matrix by some real/imaginary or in general complex quantity—such a marginal scenario should not concern us, as it has no effect on the ratio of the state vector components (which is of interest for us), and this can be directly deduced via the state vector discrete evolution equations depicted in Eq. (54).](A19)
To more efficiently study the dynamical behavior of quantum arrangements in the quasistatic limit (T → ∞, T: period of evolution), the various system parameters (e.g., gain and detuning coefficients for the -symmetric-like Hamiltonian of Sec. II B) are typically written directly as functions of the scaled time coordinate (this entails that if t and T increase multiplicatively by the same factor, then the system parameters remain unaltered). This in turn implies that matrices H, M will depend directly on , and the same will also be true for their respective eigenvalues and eigenvectors (the state vector will of course depend on both , and thus, the aforementioned multiplicative rule will not hold then, as the subsequent WKB analysis will show).
- Pertaining to -symmetric-like settings, at many instances (particularly in numerical computations), a normalization scheme with respect to the coupling constant κ is employed. This arises after recasting the Hamiltonian evolution formula in Eq. (14) of Sec. II B asThe state vector dynamics can then be studied in terms of variables . The state vector itself, its components, and their ratio remain unaffected (the only change is that now they are written as functions of rather than t). The new system and Hamiltonian matrices read as , which implies for the eigenvalues that . Moreover, regarding the previous bullet point, it applies that , and thus, the scaled time coordinate will not vary if we first normalize with respect to κ.(A20)
2. Retrieving the dynamical, geometric, and higher-order phase factors via WKB theory
To summarize, we shall provide the following concluding notes:
In the present appendix, we have shown how via the WKB method an approximate expression can be attained for the state vector component a(t) in the adiabatic limit T → ∞ [analogous formulas can also be derived for component b(t)]. A direct correlation between the two leading-order terms of the WKB series with the dynamical [ϕ±(t)] and geometric phases has been demonstrated [see Eqs. (A28) and (A32)]. The conditions for approximating a(t) as the product of the aforementioned phase factors [see Eq. (A33)] have been outlined [see Eq. (A34)], while in the same time, a difference–differential relation has been acquired that can yield all higher-order terms/phases with j ≥ 2 [see Eq. (A35) along with the general constraints for the validity of the WKB method in Eq. (A24)]. It should be highlighted that exactly the same results can be attained had we originally emphasized on the differential equation that the transformed component satisfies [see Eq. (12) of Sec. II A], after of course reinstating a(t) in the retrieved WKB formulas [see Eq. (10) of Sec. II A].
A few notes must also be made regarding the terminology. In Appendix A 1, the system matrix M (and Hamiltonian H) and its eigenvalues and eigenvectors (and the associated eigenvector component ratios) were written as functions of time t. As explained though in Appendix A 1 (and at the beginning of Sec. II A), the above quantities depend directly on the scaled time coordinate , and that is why, here, they were written as a function of (the same applies also for terms , including the geometric phase ). This comes in contrast with the state vector component a (and b), which depends on both and T as becomes readily evident from Eqs. (A21) and (A22) (the latter in essence implies that if t and T are multiplied by the same factor, a will vary despite that remains unchanged). Consequently, we write [or simply a(t)] and not , to correctly reflect the underlying functional dependencies. In a similar fashion, the dynamical phase ϕ± is a function of both and T [see Eqs. (A27) and (A28)] and thus can be written equivalently as ϕ±(t) or [in the analysis that follows, we might also use the auxiliary variable , which will depend solely on ]. To summarize, an arbitrary quantity q that is a function only of will be written alternatively as q(t) or [see for instance or λ± in Eqs. (A27) and (A28)], while if q is a function of both and T, then this will be expressed as q(t) or [see the component a in Eqs. (A21), (A22), and (A33)].
A fundamental difference between the dynamical and geometric phases can be underlined at this point: the former depends on how fast or slow the Hamiltonian varies with time (i.e., the magnitude of T), while the latter depends solely on the geometry of the path followed within parameter space (thus, the term “geometric phase”). Essentially, each point along can be described by a specific value of , which is independent from the choice of period T (this does not apply for t, as if for example T is increased, then t will signify points along that are closer to the starting point of evolution). This stems from the fact that the various Hamiltonian parameters are functions only of in our adiabatic analysis, and thus, can be fully parameterized in terms of (and not of t). Consequently, all points of will be characterized by a unique (the 1–1 correspondence is guaranteed via variable ), but not from a unique ϕ±(t) (given that the latter depends not only on but also on T). Moreover, it should be noted that given the 1–1 mapping between and , the integrals in Eqs. (A27), (A31), and (A32) can be equivalently expressed as line integrals along .
3. Asymptotic behavior of Airy functions for arguments of large absolute value
By this point, the efficacy of the WKB approach in analyzing the slowly varying dynamics of generalized Hamiltonian settings within high orders of accuracy has been clearly illustrated. In the present and following sections, we shall apply such a methodology to attain asymptotic expressions for two distinct classes of functions, namely the Airy ( Appendix A 3) and the modified Bessel ( Appendix A 4) functions.
At this point, we can make an interesting observation: the coefficient c1 changes from zero when z > 0, to a nonzero value when z < 0. So how can it be that Ai(z) is analytic everywhere, but it does not have a uniform asymptotic expansion throughout the complex plane? This famous paradox in complex analysis is known as the Stokes phenomenon and is largely attributed to the presence of multivalued functions in the asymptotic expressions of analytic functions. [Consider, for instance, and that we employ the principal branches for both (multivalued) functions z3/2 and z−1/4 (branch cut is then along axis z < 0). Taking |z| = ρ > 0, it comes that as arg(z) → π, then , while as arg(z) → −π, then . Hence, at the same point along the negative real axis (z = −ρ), attains different values and this dictates that c1 should have different values below and above axis z < 0. Such a fact has been already pointed out when deriving the asymptotic formula of Eq. (A51).]
To resolve the aforementioned inconsistency, Stokes considered what happens as z moves along a circle around the origin of the complex plane. He realized that the exponential factors in Eq. (A47) assumed both large and small absolute values intermittently: while one term is superior, the other is inferior and vice versa (this is contingent upon whether the exponent has a positive or a negative real part). Furthermore, a “jump” in the value of a coefficient (c1 or c2) must not compromise the integrity of the asymptotic approximation as a whole—if a new exponential term is set to emerge due to the switching of its coefficient from zero to nonzero, it should occur when it is the least noticeable. This led to Stokes’ ingenious observation that a jump could only manifest at points where the switching coefficient in question corresponds to an inferior term and the superior term reaches its maximum.44 (In a much later study,45 Berry actually showed that such abrupt jumps do not occur discontinuously but rather in a smooth fashion, by truncating the dominant part of the asymptotic series expansion near its smallest element and subsequently studying the subdominant contributions via Dingle’s method of Borel sums.46) These points are known to lie on the Stokes lines, which in Eq. (A47) occur when Im(z3/2) = 0 ⇔ arg(z) = 0, ±2π/3. [Note that the locations where Re(z3/2) = 0 ⇔ arg(z) = ±π/3, π define the anti-Stokes lines, and there and become comparable in magnitude.] Then, for arg(z) = 0, it is that asymptotically dominates over , while this is reversed when arg(z) = ±2π/3. Hence, if c1 (c2) exhibits a jump, this better takes place when arg(z) = ±2π/3 [arg(z) = 0]. The above rationale is of very general application to a class of functions frequently emerging in physical problems and exhibiting rapidly varying (oscillating or exponentially growing/decaying) components (examples include the Bessel, parabolic cylinder, and hypergeometric functions, or even the Fresnel and elliptic integrals). In this respect, if the leading behaviors in Eq. (A47) are given by eP(z) and eQ(z) (which are also assumed to display significantly different dynamics across the complex space), the (anti-)Stokes curves will be described by the equation Im[P(z) − Q(z)] = 0 {Re[P(z) − Q(z)] = 0}, and in their vicinity, one of the solutions will prevail (both solutions will be comparable).43
4. Asymptotic behavior of modified Bessel functions for orders of large absolute value
At this point, we should highlight that great care must be given in all the aforementioned computations, owing to the presence of multivalued functions in the various formulas. The selection of a specific branch causes branch cuts to appear in complex space, which subsequently introduces complications even in the simplest of the calculations as laboriously depicted in Sec. II B. Yet, the effect of branch cuts has not been demonstrated so far in complex integration. In Appendixes A 2– A 4, we have avoided integrating along specific paths (indefinite integrals have been used in all instances), so as not to consider scenarios where such paths intersect with branch lines. [Had this occurred, the integral in question should have been evaluated separately before and after these discontinuity curves, as we shall subsequently see, and this in turn can give rise to additional multiplication constants in the WKB expressions of Eqs. (A33), (A47), and (A59), thus modifying the already existing constants —the previously described mathematics are in essence the culprit behind the manifestation of the Stokes phenomenon (see also the detailed description of Appendix A 3)]. This simplifies the underlying analysis and enables us to better illustrate the utility of the WKB method through the examined asymptotic studies. In what follows, it will be clarified via certain examples how the path of integration affects the values of the associated complex integrals, particularly when branch lines are crossed or when the integrand displays singularities within the domain of integration.
5. A few notes on the evaluation of complex integrals in the presence of branch cuts and singularity points
Throughout the current investigation, we have employed for our calculations the principal values for the arguments ϑ of complex numbers, i.e., −π < ϑ ⩽ π (see the beginning of Sec. II B 1). Of course, the argument of a complex quantity is simply one of a garden variety of multivalued complex functions (examples are the logarithmic, the square root, the nth root with n being a positive integer, and the inverse trigonometric). In order to be mathematically consistent, we have chosen all such functions to be represented by their principal branches (for instance, , with z = ρeiϑ, ρ > 0, −π < ϑ ⩽ π), which in turn implies the emergence of branch cuts in the complex plane where discontinuities arise (e.g., for the principal square/nth root and logarithm, the branch cut extends from the origin along the negative real axis to negative infinity). In general, branch cuts designate the locations that the different sheets/branches of a multivalued function meet and thus are typically employed in complex analysis to delimit the function’s domain in regions where it becomes single-valued and well-behaved (here, such a region corresponds to the principal branch).
Now, the question comes as to what happens if a line integral crosses such a branch cut? So far, in the evaluation of the various complex integrals, the latter scenario was not taken into account, as our emphasis was on demonstrating the effectiveness of the WKB method in analyzing the asymptotic behavior of generalized Hamiltonian settings (and subsequently of particular function categories), without delving into intricate mathematical details. In what follows, we shall show via specific examples how to address such special situations, while the effect of poles/singularities will also be separately discussed.
We shall conclude the present discussion, by providing the following remarks:
- In the current analysis, we have considered the impact of using the principal branch when evaluating line integrals involving multivalued complex functions (the presented approach is general enough and can be applied if other branches are employed). Yet, special care must be taken even when performing more basic computations. For instance, the relation is not always applicable and an example of this is if z1 = −1 and z2 = i,Similarly, equality ln(z1z2) = ln(z1) + ln(z2) does not uphold, and to see this, we can set again z1 = −1 and z2 = i,
More extensive operations of this type (especially when it comes to square and nth roots) can be found in Sec. II B. In general, to avoid any inconsistencies, all calculations must be done step by step to ensure that each time the principal branches of the different multivalued functions are employed.
Clearly, all integrals evaluated here are well-defined, since they result in finite numerical values, which are also independent of the chosen parameterization. However, in general, for the contour integral ∫Cf(z)dz to be properly defined, it is necessary and sufficient that C is a piecewise smooth curve and f(z) is piecewise continuous along C. Both of the latter conditions are satisfied in all examined cases.
- Certain benefits of employing the mode notation, which was introduced in Sec. III, can become apparent. To see this, let us consider the evaluation of the square roots contained in the eigenvalue formulas of Eq. (A3) for a two-mode Hamiltonian arrangement [an analogous rationale can apply for n-mode Hamiltonians, where nth (and lower-order) roots emerge in the corresponding eigenvalue expressions]. When the phase of the complex quantity beneath the square root shifts from the second to the third quadrant (π− → −π+) or vice versa (−π+ → π−), a discontinuity is anticipated in the evolution of λ±(t) within the complex eigenvalue space given that {z = ρeiϑ, ρ > 0, with ϑ, ϑo ∈ (−π, π]},
Indeed, in the special case of a -symmetric-like system where , such an abrupt transition takes place when λ±(t) cross the real axis as confirmed from Fig. 2. Meanwhile, when using the representation, the trajectories that follow remain continuous in any circumstance (see Figs. 2 and 4–6) and this is of course attributed to the definition of the notation [i.e., it reorders modes (±) at each time step so as to attain a smooth eigenspectra evolution]. The usefulness then of such a mode formalism is multifold: (i) it helps us to better visualize the evolution of both eigenvalues (e.g., see Fig. 2) and eigenvectors (e.g., see Fig. 3 depicting the eigenvector component ratios), as abrupt “jumps” will not take place in the respective diagrams, (ii) it facilitates the analysis at many points, since it leads to well-defined derivatives and to more manageable integrals, where we do not have to worry about the emergence of branch cuts and where the various theorems/lemmas of complex analysis can become directly applicable, (iii) it yields physically consistent results, as the eigenspectra of an actual physical setting can never exhibit a discontinuous behavior (the eigenstates should vary in a continuous fashion, and the same applies for the associated mode population dynamics, or, more specifically, for the projections of the state vector on the system eigenvectors, assuming of course that the state vector also varies continuously).
APPENDIX B: RIGOROUS ANALYTICAL TREATMENT OF THE MODE POPULATION DYNAMICS
1. Deriving the mode population equations for generalized Hamiltonian settings
At this point, we shall retrieve the mode population dynamics for two-level Hamiltonian arrangements. We will be using the scaled time coordinate (=t/T), so as (i) the dependencies of the different variables and parameters from and T can be clearly depicted and (ii) to facilitate the slowly varying study in Appendix B 3, where T serves as a tuning knob to achieve adiabaticity. Moreover, the mode formalism shall be employed in order to avoid potential discontinuities that may affect the evaluation of complex integrals and, additionally, lead to ill-defined derivatives (see the relevant discussion in Appendix A 5).