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

Quantum transport in nonlinear Rudner-Levitov models

Lei Du Beijing Computational Science Research Center, Beijing 100193, China    Jin-Hui Wu Center for Quantum Sciences and School of Physics, Northeast Normal University, Changchun 130024, China    M. Artoni [email protected] Department of Engineering and Information Technology and INO-CNR Sensor Lab, Brescia University, 25133 Brescia, Italy    G. C. La Rocca [email protected] NEST, Scuola Normale Superiore, 56126 Pisa, Italy
Abstract

Quantum transport in a class of nonlinear extensions of the Rudner-Levitov model is numerically studied in this paper. We show that the quantization of the mean displacement, which embodies the quantum coherence and the topological characteristics of the model, is markedly modified by nonlinearities. Peculiar effects such as a “trivial-nontrivial” transition and unidirectional long-range quantum transport are observed. These phenomena can be understood on the basis of the dynamic behavior of the effective hopping terms, which are time and position dependent, containing contributions of both the linear and nonlinear couplings.d nonlinear couplings.

I Introduction

The Rudner-Levitov (RL) model RLmodel , which can be viewed as a simple non-Hermitian extension of the Su-Schrieffer-Heeger (SSH) model SSH , has been predicted to support quantized quantum transport with topological protection RLmodel ; RLarxiv has been verified experimentally in a lattice of optical waveguides RL2015 . This important feature not only shows the possibility of accessing nontrivial topological phases in physical systems where dissipations are intrinsic and unavoidable, but also provides a direct way to detect the topology of such systems by monitoring their bulk properties. Specifically, for a dimerized ladder of discrete quantum states, with one sublattice lossy and the other one neutral, the mean displacement of a particle which is initialized at a neutral site shows well quantized behaviors, depending on the ratio between the intra- and inter-cell coupling constants.

Recently, the intersection of nonlinear optics and topological photonics has unlocked a series of peculiar effects that have no counterparts in condensed matter topologies NonTopoRev , such as magnet-free nonreciprocity MFnr1 ; MFnr2 , active tunability tuna1 ; tuna2 ; tuna3 , self-induced topological transitions Hadad ; Hadad2 ; Ezawa , topological solitons tsoliton1 ; tsoliton2 ; tsoliton3 ; tsoliton4 ; chong2021 , and topological frequency conversions FC1 ; FC2 ; FC3 . Most recently, Xia et al. have considered a nonlinear PT-symmetric optical waveguide lattice to investigate the mutual interplay between topology, PT symmetry, and nonlinearity XiaScience , demonstrating that global effects, i.e., topological and PT transitions, can be actively controlled by the local nonlinearity. The RL model has already been generalized by including on-site Kerr-type nonlinearities, which shows that the quantized quantum transport is severely affected by the nonlinearity induced decoherence effects Rapedius and that the nonlinear effect becomes pronounced in the PT-broken region RLSR . However, the study of the influence of nonlinearity on the quantum transport in the RL model is still at the initial stage. In particular, work on how nonlinearities affect the hopping terms, rather than the on-site energies, is still missing; the latter in fact may affect quantum transport in the RL model precisely because its topological properties are dictated by the ratio between the intra- and inter-cell coupling constants.

In this paper, we consider specific extensions of the RL model with different types of nonlinear couplings and numerically study the quantum transport therein. As it happens for the case of on-site nonlinearities Rapedius ; RLSR , our results show that the nonlinear coupling generates departures from the typical quantized behavior of the particle transport. Our hopping nonlinearities turn out to be responsible for new effects, such as (i.) a “trivial-nontrivial” transition, by which the mean displacement will always take the value of 0 or 11 regardless of the linear coupling imbalance, and (ii.) a long-range transport, by which the particle can move unidirectionally across many unit cells before its decay. These phenomena can be understood on the basis of the dynamic behavior of the effective time and position dependent hopping terms containing contributions of both the linear and nonlinear couplings. Our findings are relevant to experimental platforms, such as arrays of coupled optical waveguides and cavities, but may also be adapted to acoustic or electric circuit resonators as well as superconducting quantum circuits Kono ; Smirn ; Xue .

II The Rudner-Levitov model

Refer to caption
Figure 1: (Color online) Schematic illustration of the Ruder-Levitov model, where μ\mu (ν\nu) and γa\gamma_{a} are the intracell (intercell) coupling constant and the decay rate of the lossy sites, respectively.

We start by briefly reviewing the original RL model RLmodel ; RLarxiv ; RL2015 , where each unit cell of the one-dimensional dimerized lattice contains a lossy site AA and a neutral site BB, as shown in Fig. 1. The intra- and inter-cell coupling constants are denoted by μ\mu and ν\nu, respectively, while the decay rate of each lossy site is denoted by γa\gamma_{a}. In the lossless limit γa0\gamma_{a}\rightarrow 0, such a linear bipartite lattice is equivalent to the SSH model SSH which shows a topological transition depending on the ratio μ/ν\mu/\nu zak ; asboth . It has been shown that the nonvanishing loss in the RL model plays the key role for observing quantized quantum transport. The real-space Hamiltonian of the RL model is given by (=1\hbar=1 hereafter)

H0=m[iγa|m,am,a|(ν|m,am1,b|μ|m,am,b|+H.c.)],\begin{split}H_{0}&=\sum_{m}\Big{[}-i\gamma_{a}|m,a\rangle\langle m,a|-(\nu|m,a\rangle\langle m-1,b|\\ &\quad\,-\mu|m,a\rangle\langle m,b|+\text{H.c.})\Big{]},\end{split} (1)

where m[N,N]m\in[-N,N] labels the unit cells of the lattice; |m,a|m,a\rangle and |m,b|m,b\rangle denote the Wannier states of sites AA and BB in the mmth unit cell, respectively.

Setting |ψ(t)=m[am(t)|m,a+bm(t)|m,b]|\psi(t)\rangle=\sum_{m}[a_{m}(t)|m,a\rangle+b_{m}(t)|m,b\rangle] as a general eigenstate of the system, with ama_{m} (bmb_{m}) the field amplitude of AA (BB) in the mmth unit cell, it is straightforward to attain the following dynamic equations

amt=γaam+iνbm1+iμbm,bmt=iμam+iνam+1.\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i\nu b_{m-1}+i\mu b_{m},\\ \frac{\partial b_{m}}{\partial t}&=i\mu a_{m}+i\nu a_{m+1}.\end{split} (2)

Hereafter in this paper, we assume μ=0.5δg\mu=0.5-\delta g and ν=0.5+δg\nu=0.5+\delta g with δg\delta g denoting the coupling imbalance (|δg|0.5|\delta g|\leq 0.5). Then, one can numerically calculate the mean displacement of a particle before its decay, i.e.,

Δm=mm02γa|am(t)|2𝑑t.\langle\Delta m\rangle=\sum_{m}m\int_{0}^{\infty}2\gamma_{a}|a_{m}(t)|^{2}dt. (3)

This is equivalent to the winding number of the relative phase between components of the Bloch wave function [i.e., the eigenstate of the Fourier transformation of H(m)H(m)] and is thus topologically protected RLmodel ; RLarxiv ; zak ; asboth . Specifically, if one initializes a single particle at a neutral site, the mean displacement Δm\langle\Delta m\rangle takes the value of 0 for ν<μ\nu<\mu and takes the value of 11 for ν>μ\nu>\mu, which precisely predicts the topological phase transition of this model.

Such an archetype model has been well investigated both theoretically RLmodel ; RLarxiv and experimentally RL2015 , and has also been extended in many directions Rapedius ; RLSR ; LiExtend ; LieuExtend ; LonghiExtend ; JinExtend ; YuceExtend ; RLdl ; SAP ; WuExtend ; BergExtend . In the following, we generalize the RL model to a few instances with nonlinear hopping terms and study the quantum transport therein.

III Nonlinearity induced trivial-nontrivial transition

We anticipate that all our nonlinear extensions have symmetric nonlinear couplings so that the norm of a quantum state still evolves according to dψ|ψ/dt=m2γa|am|2d\langle\psi|\psi\rangle/dt=-\sum_{m}2\gamma_{a}|a_{m}|^{2} as in the linear RL model (see Appendix A); this validates in turn the definition of the mean displacement in Eq. (3). In the absence of losses, this assures the conservation of the standard norm of a quantum state; other nonlinear models, such as those related to the Ablowitz-Ladik system NDDE ; ALpra ; renormAL , do not have this property and are outside the scope of this paper.

Refer to caption
Figure 2: (Color online) Mean displacement Δm\langle\Delta m\rangle of the particle versus coupling imbalance δg\delta g for (a) model AA and (b) model BB, with initial state |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle, integral time γaT=50\gamma_{a}T=50, and different values of nonlinear coefficient UU. (c) Δm\langle\Delta m\rangle versus δg\delta g for model AA with |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle, U=0U=0, and different values of integral time TT. (d) Δm\langle\Delta m\rangle versus δg\delta g for model AA, with identical parameters to those in (a) except for μμ\mu\rightarrow-\mu and νν\nu\rightarrow-\nu. Other parameters are μ=0.5δg\mu=0.5-\delta g, ν=0.5+δg\nu=0.5+\delta g andγa=2\gamma_{a}=2.

Model A. First, consider the case in which the intercell coupling terms contain a Kerr-type nonlinear part that depends on the field amplitudes of both lossy and neutral sites. The relevant dynamics follows the nonlinear Schrödinger equations

amt=γaam+iμbm+i(νξm)bm1,bmt=iμam+i(νξm+1)am+1.\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i\mu b_{m}+i(\nu-\xi_{m})b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i\mu a_{m}+i(\nu-\xi_{m+1})a_{m+1}.\end{split} (4)

where ξm=U(|am|2+|bm1|2)\xi_{m}=U(|a_{m}|^{2}+|b_{m-1}|^{2}) corresponds to the nonlinear part of the intercell coupling between the mmth and (m1)(m-1)th unit cells, the strength of which is described by the nonlinear coefficient UU. Eq. (4) embeds a non-Hermitian extension of the model in Ref. Hadad and can be employed to describe arrays of coupled optical and acoustic cavities, as well as circuit resonators.

We plot in Fig. 2(a) the mean displacement Δm\langle\Delta m\rangle as a function of the coupling imbalance δg\delta g for model AA with the initial state |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle. In the absence of nonlinearities (U=0U=0), the mean displacement of the particle before its decay is quantized with two integer values 0 and 11: the particle initially localized at the central neutral site will hop to the lossy site in the right (m=1m=1) unit cell and then decays from the lattice if ν>μ\nu>\mu, or it will hop to the left lossy site within the initial (m=0m=0) unit cell and then decays if ν<μ\nu<\mu. Note that the transition from Δm=0\langle\Delta m\rangle=0 to Δm=1\langle\Delta m\rangle=1 here is not perfectly quantized due to the finite integration time. As shown in Fig. 2(c), the quantization behavior becomes more and more ideal with the increase of the integration time TT. As the nonlinear coefficient increases from zero, the mean displacement deviates from the quantized behavior gradually. In particular, the original “topological trivial” region (where Δm=0\langle\Delta m\rangle=0 in the linear case RLmodel ; RLarxiv ) almost disappears upon increasing UU, with the values of Δm\langle\Delta m\rangle approaching unity over the whole parametric range. In other words, the particle always hops to the right unit cell for large enough UU, regardless of the relative values of μ\mu and ν\nu.

Model B. We consider next the case in which the intracell coupling is instead modified by a Kerr-type nonlinearity. The dynamics can now be described by,

amt=γaam+i(μζm)bm+iνbm1,bmt=i(μζm)am+iνam+1,\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i(\mu-\zeta_{m})b_{m}+i\nu b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i(\mu-\zeta_{m})a_{m}+i\nu a_{m+1},\end{split} (5)

with ζm=U(|am|2+|bm|2)\zeta_{m}=U(|a_{m}|^{2}+|b_{m}|^{2}) and the corresponding numerical results are plotted in Fig. 2(b). It can be seen that the mean displacement for model BB shows exactly reverse behaviors compared with those for model AA: upon increasing UU, the originally “nontrivial” region (where Δm=1\langle\Delta m\rangle=1 in the linear case) almost disappears. In this case, the particle always hops from a neutral site to the left lossy one within the same unit cell.

Refer to caption
Figure 3: (Color online) Dynamic evolutions of displacement Δm(t)\Delta m(t) (upper) and effective coupling contrast ZmZ_{m} (lower) for model AA, with (a) U=0.5U=0.5, (b) U=3U=3, and (c) U=5U=5. (d) Dynamic evolutions of effective coupling contrast Z1Z_{1} with different values of UU. The initial state is assumed as |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle. Other parameters are δg=0.4\delta g=-0.4 for (a)-(c), δg=0.1\delta g=-0.1 for (d), and γa=2\gamma_{a}=2.

The results above can be understood from the effective inter- and intra-cell couplings which contain the contributions of the nonlinear parts. As an example, we define for model AA the effective coupling contrast Zm=|νeff,m||μ|Z_{m}=|\nu_{\text{eff},m}|-|\mu| with νeff,m=νξm\nu_{\text{eff},m}=\nu-\xi_{m} the effective intercell coupling constant, and plot in Fig. 3 the dynamic evolutions of ZmZ_{m} with different values of UU. Moreover, we also plot the evolutions of the time-dependent particle displacement Δm(t)=mm0t2γa|am(t)|2𝑑t\Delta m(t)=\sum_{m}m\int_{0}^{t}2\gamma_{a}|a_{m}(t^{\prime})|^{2}dt^{\prime} Rapedius showing how it reaches its final value. In the absence of nonlinearities, the coupling contrast ZmZ_{m} reduces to Z=|ν||μ|Z=|\nu|-|\mu|, which is position independent and can be used to predict the transition of Δm\langle\Delta m\rangle, i.e., Δm=1\langle\Delta m\rangle=1 if Z>0Z>0 and Δm=0\langle\Delta m\rangle=0 if Z<0Z<0. In the presence of nonlinearities, however, ZmZ_{m} becomes dependent on mm. It can be seen from Figs. 3(a)-3(c) that for δg=0.4\delta g=-0.4 (ν=0.1\nu=0.1), the value of Zm=1Z_{m=1} increases with UU from negative to positive monotonously. On the other hand, for δg=0.1\delta g=-0.1 (ν=0.4\nu=0.4) as shown in Fig. 3(d), with the increase of UU, the value of Zm=1Z_{m=1} first diminishes until Zm=1=|μ|Z_{m=1}=-|\mu|, then it turns to increase with UU gradually from negative to positive. This is also why the the mean displacement shows a non-monotonic behavior as UU increases in this case. In Appendix B, we provide a detailed explanation for the physical meaning of the positive and negative fractional values of Δm\langle\Delta m\rangle. As a consequence, the topology in the “trivial” region can be modified by the nonlinearities and thus one can always observe nontrivial quantum transport for large enough UU. Although the region where the contrast ZmZ_{m} is modified by the nonlinearity disappears with time, its duration is long enough for the particle displacement to reach the final values [see the figures in the upper row of Figs. 3(a)-3(c)]. Intuitively, the influence of the nonlinearities disappears after the particle spreads out and decays from the lattice via the lossy sites. Similarly, the “nontrivial-to-trivial” transition of model BB can also be understood from the effective coupling contrast.

In Eqs. (4) and (5), we have assumed that the linear and nonlinear parts of the coupling terms tend to cancel each other, i.e., the signs of μ\mu and ν\nu are opposite to that of UU. Such an assumption is responsible for the non-monotonic behavior of Δm\langle\Delta m\rangle as UU changes. For comparison, we reverse the signs of μ\mu and ν\nu (i.e., μμ\mu\rightarrow-\mu and νν\nu\rightarrow-\nu) in Eq. (4) and plot in Fig. 2(d) Δm\langle\Delta m\rangle versus δg\delta g in this case. It can be seen that the values of Δm\langle\Delta m\rangle in the originally trivial region increase monotonously with UU to approach unity over the whole parametric range. This is different from the non-monotonic behavior in Fig. 2(a). Such a result is due to the fact that the effective coupling contrast Zm=1Z_{m=1} increases with UU monotonously in this case.

At the end of this section, we would like to point out that the “trivial-nontrivial” transition can also be observed even if the nonlinear coupling terms only depend on the field amplitudes of the neutral sites. To prove this, we consider in Appendix C such a case and demonstrate the behaviors of the quantum transport therein.

IV Nonlinearity induced unidirectional long-range displacement

Refer to caption
Figure 4: (Color online) Mean displacement Δm\langle\Delta m\rangle of the particle versus coupling imbalance δg\delta g for [(a) and (b)] model CC and (c) model DD, with initial state |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle and different values of nonlinear coefficient UU. We assume γa=0.2\gamma_{a}=0.2 for (a) and (c), while γa=2\gamma_{a}=2 for (b).

Model {C. & D.} Finally, we consider in this section the case in which the nonlinear coupling constants are only dependent on the field amplitudes of the lossy sites. The relevant dynamic equations take the form (Model C.),

amt=γaam+iμbm+i(νχm)bm1,bmt=iμam+i(νχm+1)am+1\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i\mu b_{m}+i(\nu-\chi_{m})b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i\mu a_{m}+i(\nu-\chi_{m+1})a_{m+1}\end{split} (6)

with χm=U|am|2\chi_{m}=U|a_{m}|^{2} describing the nonlinear part of the intercell couplings in this case. Similarly, when the intracell couplings are modified by the field amplitudes of the lossy sites in a nonlinear manner one has for the dynamic equations (Model D.D.),

amt=γaam+i(μχm)bm+iνbm1,bmt=i(μχm)am+iνam+1.\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i(\mu-\chi_{m})b_{m}+i\nu b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i(\mu-\chi_{m})a_{m}+i\nu a_{m+1}.\end{split} (7)

As in the last section, we first examine the mean displacement Δm\langle\Delta m\rangle of the particle versus the coupling imbalance δg\delta g for both models CC and DD, as shown in Figs. 4(a)-4(c). Interestingly, we find that long-range particle displacements, i.e., Δm<0\langle\Delta m\rangle<0 or Δm>1\langle\Delta m\rangle>1, are allowed within these two extensions of the RL model. As shown in Fig. 4(a), for instance, the value of Δm\langle\Delta m\rangle can be smaller than 1-1 for large UU, implying that the particle initially located at a neutral site can hop on average several times to the left before its decay if the nonlinearities are strong enough. Note that a relatively small γa\gamma_{a} (γa=0.2\gamma_{a}=0.2) is assumed in this case in order to observe the such unidirectional long-range displacements, while for large γa\gamma_{a} the mean displacement becomes less sensitive to UU, as shown in Fig. 4(b). Moreover, model DD displays reverse behaviors of the mean displacement, i.e., the particle will move to the right with Δm>1\langle\Delta m\rangle>1 for large enough UU.

Refer to caption
Figure 5: (Color online) Dynamic evolutions of displacement Δm(t)\Delta m(t) (upper) and effective coupling contrast ZmZ_{m} (lower) for model CC, with (a) U=0.5U=0.5, (b) U=3U=3, and (c) U=5U=5. The initial state is assumed as |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle. Other parameters are δg=0.4\delta g=-0.4 and γa=0.2\gamma_{a}=0.2.

The underlying physics of the above results can be qualitatively understood as follows. For positive δg\delta g the intercell coupling is stronger than the intracell one, thus the particle initially localized at a neutral site will hop to the right lossy site and eventually decay. In this case, the value of the mean displacement cannot be larger than 11, because the particle will not keep on moving due to the weak intracell coupling. In contrast, for negative δg\delta g, the particle will hop to the left lossy site within the initial unit cell (i.e., the 0th unit cell in Fig. 4). For model CC, the effective intercell coupling between the 0th and 1-1th unit cells is enhanced after the particle has moved to the site |m=0,a|m=0,a\rangle. If the nonlinearity is strong enough, the particle can keep on moving towards the left and thereby the mean displacement should increase with UU. Of course, owing to the dissipation at each lossy site, the mean displacement is still finite since the effective intercell coupling will no longer be stronger than the adjacent intracell coupling after the particle has moved across several lossy sites. This is also why the long-range particle displacement tends to vanish if the decay rate γa\gamma_{a} is large enough. The reverse behaviors in model DD can be interpreted in a similar way.

Similar to the last section, we plot in Figs. 5(a)-5(c) the dynamic evolutions of the time-dependent displacement Δm(t)\Delta m(t) as well as the effective coupling contrast Zm=|νχm||μ|Z_{m}=|\nu-\chi_{m}|-|\mu| of model CC with different values of UU. Once again, one can find that the values of ZmZ_{m} in a certain region increase with UU. Unlikely in Fig. 3, however, where only the nearest-neighbor intercell coupling (the coupling between the 0th and the 11th unit cells) is markedly modified by the nonlinearities, here model CC exhibits a larger modification region that extends toward the left side of the model. This again reflects the origin of the long-range displacements. For all cases in Fig. 5, the particle displacement reaches its final value before the modification region disappears.

Refer to caption
Figure 6: (Color online) Dynamic evolutions of initial state |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle for model CC with (a) U=5U=5 and δg=0.4\delta g=-0.4, (b) U=3U=3 and δg=0.4\delta g=-0.4, (c) U=0.5U=0.5 and δg=0.4\delta g=-0.4, and (d) U=0.5U=0.5 and δg=0\delta g=0. In all cases, we take γa=0.2\gamma_{a}=0.2.

The results in Fig. 5 can also be verified by the time evolutions of the initial state |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle. As shown in Figs. 6(a)-6(c), the particle transport tends to be unidirectional as the nonlinearities strengthen. Note that for δg=0\delta g=0, the lattice is homogeneous at the initial time, such that the particle transport should be almost symmetric, insensitive to the value of UU. As an example, Fig. 6(d) shows the dynamic evolution of the initial state for δg=0\delta g=0 and U=0.5U=0.5, which is similar to the symmetric diffusion in a one-dimensional homogeneous lattice. In Fig. 6(c), however, the excitation is confined within a narrow region around m=0m=0 because the effective intercell couplings are extremely weak (both the linear and nonlinear parts are weak) in this case.

V Conclusions

In summary, we have extended the Rudner-Levitov (RL) model allowing for nonlinear hopping and assessed how nonlinearities affect the mean displacement in quantum transport. At variance with on-site nonlinearities, dynamic modifications of the ratio between the effective intra- and inter-cell couplings directly affect the behavior of quantum transport related to the topological characteristics of the RL model. Our results encompass different extensions of the RL model, which are interpreted qualitatively in terms of the local and time-dependent effective coupling contrast function Zm(t)Z_{m}(t). There is one specific extension (See Sect. IV) whereby for strong nonlinearities and low losses unidirectional long-range quantum transport may emerge. The present work is relevant to experimental platforms that hinge on tight-binding Hamiltonians, including non-Hermitian Hamiltonians and those describing nonlinear effects in optics, in acoustics, and in electronics.

Acknowledgments

This work is supported by the National Natural Science Foundation of China (No. U1930402 and No. 12074061), the Cooperative Program by the Italian Ministry of Foreign Affairs and International Cooperation (MAECI) through Italy-China International Cooperation (PGR00960), and the “Laboratori Congiunti” 2019 Program of National Research Council (CNR) of Italy.

Appendix A Norm evolution with symmetric couplings

It has been shown that the norm of a quantum state should evolve according to

ddtψ|ψ=m2γa|am|2\frac{d}{dt}\langle\psi|\psi\rangle=-\sum_{m}2\gamma_{a}|a_{m}|^{2} (8)

to guarantee the validity of the definition of the mean displacement in Eq. (3RLmodel . In this appendix, we would like to prove that Eq. (8) still holds for nonlinear RL models with symmetric couplings. For example, according to Eq. (4) in the main text, the norm of a quantum state in model AA should follows

dψ|ψdt=(ddtψ|)|ψ+ψ|(ddt|ψ)=m[(damdt)am+(dbmdt)bm+am(damdt)+bm(dbmdt)]=m[γa|am|2igrbmami(glξm)bm1am+γa|am|2+igrbmam+i(glξm)bm1amigrambmi(glξm+1)am+1bm+igrambm+i(glξm+1)am+1bm]=m2γa|am|2.\begin{split}\frac{d\langle\psi|\psi\rangle}{dt}&=(\frac{d}{dt}\langle\psi|)|\psi\rangle+\langle\psi|(\frac{d}{dt}|\psi\rangle)\\ &=\sum_{m}\Big{[}(\frac{da_{m}^{*}}{dt})a_{m}+(\frac{db_{m}^{*}}{dt})b_{m}\\ &\quad\,+a_{m}^{*}(\frac{da_{m}}{dt})+b_{m}^{*}(\frac{db_{m}}{dt})\Big{]}\\ &=\sum_{m}\Big{[}-\gamma_{a}|a_{m}|^{2}-ig_{r}b_{m}^{*}a_{m}-i(g_{l}-\xi_{m})b_{m-1}^{*}a_{m}\\ &\quad\,+\gamma_{a}|a_{m}|^{2}+ig_{r}b_{m}a_{m}^{*}+i(g_{l}-\xi_{m})b_{m-1}a_{m}^{*}\\ &\quad\,-ig_{r}a_{m}^{*}b_{m}-i(g_{l}-\xi_{m+1})a_{m+1}^{*}b_{m}\\ &\quad\,+ig_{r}a_{m}b_{m}^{*}+i(g_{l}-\xi_{m+1})a_{m+1}b_{m}^{*}\Big{]}\\ &=-\sum_{m}2\gamma_{a}|a_{m}|^{2}.\end{split} (9)

Similarly, one can verify that models B, C, and D also respect the norm evolution in Eq. (8) due to the symmetric couplings.

On the other hand, the RL model could also be extended to include asymmetric couplings. As an example, one could consider a model described by the dynamic equations

amt=γaam+i(μχa,m)bm+i(νχa,m)bm1,bmt=i(μχb,m)am+i(νχb,m)am+1\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i(\mu-\chi_{a,m})b_{m}+i(\nu-\chi_{a,m})b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i(\mu-\chi_{b,m})a_{m}+i(\nu-\chi_{b,m})a_{m+1}\end{split} (10)

with χα,m=U|αm|2\chi_{\alpha,m}=U|\alpha_{m}|^{2} (α=a,b\alpha=a,\,b), which might be viewed as a simple dimerization of the Ablowitz-Ladik equation NDDE ; ALpra ; renormAL . For such a model, however, the evolution of the norm cannot be simply described by a sum over local decays from each lossy site and the definition of the mean displacement given in Eq. (3) cannot be applied.

Appendix B Coherent vs incoherent transport

It has been shown that the mean displacement Δm\langle\Delta m\rangle displays continuous dependence on the relative values of ν\nu and μ\mu in the case of incoherent hopping and in the absence of nonlinearities RLmodel . Moreover, similar incoherent transport behavior has been obtained with on-site nonlinearities due to the interaction-induced decoherence effect Rapedius . The underlying physics of such results is quite useful for understanding the positive and negative fractional values of Δm\langle\Delta m\rangle in Fig. 2.

For incoherent hopping, the original RL model exhibits continuous variation of Δm\langle\Delta m\rangle from 0 to 11 and thus can also take fractional values. This arises from the competition between the probabilities of the particle hopping from the initial site to all other sites. The particle initially located at site |m=0,b|m=0,b\rangle can hop to either |m=0,a|m=0,a\rangle or |m=1,a|m=1,a\rangle, corresponding to Δm=0\langle\Delta m\rangle=0 and Δm=1\langle\Delta m\rangle=1, respectively. However, the contributions of all subsequent hoppings should cancel out in this case (the hopping probabilities from |m=0,b|m=0,b\rangle to |m=1,b|m=-1,b\rangle and to |m=+1,b|m=+1,b\rangle are the same). In view of this, the value of Δm\langle\Delta m\rangle is determined by the probability of hopping to A1A_{1} initially, which is given by Δm=ν2/(ν2+μ2)\langle\Delta m\rangle=\nu^{2}/(\nu^{2}+\mu^{2}) RLmodel ; Rapedius .

For models AA and BB here considered, the fractional values of Δm\langle\Delta m\rangle can be understood in a similar way. However, the most significant difference compared with the linear model above is that the contributions of long-range hopping processes no longer exactly cancel out due to the mm-dependent effective couplings, i.e., the effective couplings on the left and right sides of the initial site are no longer symmetric. In particular, also negative fractional values of Δm\langle\Delta m\rangle are allowed, if the left-hand effective couplings are stronger than the right-hand ones.

Appendix C Nonlinear couplings with only neutral-site contributions

Refer to caption
Figure 7: (Color online) (a) Mean displacement Δm\langle\Delta m\rangle of the particle versus coupling imbalance δg\delta g for model EE with different values of UU. (b) Comparison of Δm\langle\Delta m\rangle versus δg\delta g for models AA and EE with relatively strong nonlinearities. The initial state is assumed as |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle. Other parameters are μ=0.5δg\mu=0.5-\delta g, ν=0.5+δg\nu=0.5+\delta g and γa=2\gamma_{a}=2.

In this appendix, we consider an extended RL model (model EE) where the nonlinear intercell couplings only depend on the neutral-site field amplitudes, described by the dynamic equations

amt=γaam+iμbm+i(νηm1)bm1,bmt=iμam+i(νηm)am+1\begin{split}\frac{\partial a_{m}}{\partial t}&=-\gamma_{a}a_{m}+i\mu b_{m}+i(\nu-\eta_{m-1})b_{m-1},\\ \frac{\partial b_{m}}{\partial t}&=i\mu a_{m}+i(\nu-\eta_{m})a_{m+1}\end{split} (11)

with ηm=U|bm|2\eta_{m}=U|b_{m}|^{2} in this case.

Due to the fact that the particle is initially located at a neutral site with |ψ(t=0)=|m=0,b|\psi(t=0)\rangle=|m=0,b\rangle, the nearest-neighbor intercell coupling can still be markedly modified at the initial time by the local field amplitude b0b_{0}, similar to the case of model AA. In view of this, Fig. 7(a) shows qualitatively the same behaviors as those in Fig. 2(a): as UU increases, the curve of Δm\langle\Delta m\rangle versus δg\delta g tends to be flat, with the values approaching unity over the whole parametric range. In other words, the “trivial-nontrivial” transition in model AA can still be achieved even if the nonlinear intercell couplings only depend on the neutral-site field amplitudes. As shown in Fig. 7(b), however, the effective intercell couplings are a bit weaker due to the absence of the contributions of the lossy sites, such that the values of Δm\langle\Delta m\rangle for model EE are smaller than those for model AA, especially in the region of δg<0\delta g<0. To achieve similar modification effects, stronger nonlinearities are required in this case (see for instance the red dashed and green dotted lines).

References

  • (1) M. S. Rudner and L. S. Levitov, Topological transition in a non-Hermitian quantum walk, Phys. Rev. Lett. 102, 065703 (2009).
  • (2) W. P. Su, J. R. Schrieffer, and A. J. Heeger, Solitons in polyacetylene, Phys. Rev. Lett. 42 1698-1701 (1979).
  • (3) M. S. Rudner, M. Levin, and L. S. Levitov, Survival, decay, and topological protection in non-Hermitian quantum transport, arXiv:1605.07652.
  • (4) J. M. Zeuner, M. C. Rechtsman, Y. Plotnik, Y. Lumer, S. Nolte, M. S. Rudner, M. Segev, and A. Szameit, Observation of a Topological Transition in the Bulk of a Non-Hermitian System, Phys. Rev. Lett. 115, 040402 (2015).
  • (5) D. Smirnova, D. Leykam, Y. Chong, and Y. Kivshar, Nonlinear topological photonics, Appl. Phys. Rev. 7, 021306 (2020).
  • (6) X. Zhou, Y. Wang, D. Leykam, and Y. D. Chong, Optical isolation with nonlinear topological photonics, New J. Phys. 19, 095002 (2017).
  • (7) W. Chen, D. Leykam, Y. Chong, and L. Yang, Nonreciprocity in synthetic photonic materials with nonlinearity, MRS Bull. 43, 443 (2018).
  • (8) D. Leykam, S. Mittal, M. Hafezi, and Y. D. Chong, Reconfigurable Topological Phases in Next-Nearest-Neighbor Coupled Resonator Lattices, Phys. Rev. Lett. 121, 023901 (2018).
  • (9) A. Amo, When quantum optics meets topology, Science 359, 638 (2018).
  • (10) D. A. Dobrykh, A. V. Yulin, A. P. Slobozhanyuk, A. N. Poddubny, and Y. S. Kivshar, Nonlinear Control of Electromagnetic Topological Edge States, Phys. Rev. Lett. 121, 163901 (2018).
  • (11) Y. Hadad, A. B. Khanikaev, and A. Alù, Self-induced topological transitions and edge states supported by nonlinear staggered potentials, Phys. Rev. B 93, 155112 (2016).
  • (12) Y. Hadad, J. C. Soric, A. B. Khanikaev, and A. Alù, Self-induced topological protection in nonlinear circuit arrays, Nat. Electronics 1, 178 (2018).
  • (13) M. Ezawa, Nonlinearity-induced transition in the nonlinear Su-Schrieffer-Heeger model and a nonlinear higher-order topological system, Phys. Rev. B 104, 235420 (2021).
  • (14) Y. Lumer, M. C. Rechtsman, Y. Plotnik, and M. Segev, Instability of bosonic topological edge states in the presence of interactions, Phys. Rev. A 94, 021801 (2016).
  • (15) D. Solnyshkov, O. Bleu, B. Teklu, and G. Malpuech, Chirality of topological gap solitons in bosonic dimer chains, Phys. Rev. Lett. 118, 023901 (2017).
  • (16) D. A. Smirnova, L. A. Smirnov, D. Leykam, and Y. S. Kivshar, Topological edge states and gap solitons in the nonlinear Dirac model, Laser Photonics Rev. 13, 1900223 (2019).
  • (17) W. Zhang, X. Chen, Y. V. Kartashov, V. V. Konotop, and F. Ye, Coupling of edge states and topological Bragg solitons, Phys. Rev. Lett. 123, 254103 (2019).
  • (18) L.-J. Lang, S.-L. Zhu, and Y. D. Chong, Non-Hermitian topological end breathers, Phys. Rev. B 104, L020303 (2021).
  • (19) S. Kruk, A. Poddubny, D. Smirnova, L. Wang, A. Slobozhanyuk, A. Shorokhov, I. Kravchenko, B. Luther-Davies, and Y. Kivshar, Nonlinear light generation in topological nanostructures, Nat. Nanotechnol. 14, 126 (2019).
  • (20) Y. Wang, L.-J. Lang, C. H. Lee, B. Zhang, and Y. D. Chong, Topologically enhanced harmonic generation in a nonlinear transmission line metamaterial, Nat. Commun. 10, 1102 (2019).
  • (21) D. Smirnova, S. Kruk, D. Leykam, E. Melik-Gaykazyan, D.-Y. Choi, and Y. Kivshar, Third-harmonic generation in photonic topological metasurfaces, Phys. Rev. Lett. 123, 103901 (2019).
  • (22) S. Xia, D. Kaltsas, D. Song, I. Komis, J. Xu, A. Szameit, H. Buljan, K. G. Makris, Z. Chen, Nonlinear tuning of PT symmetry and non-Hermitian topological states, Science 372, 72 (2021).
  • (23) K. Rapedius and H. J. Korsch, Interaction-induced decoherence in non-Hermitian quantum walks of ultracold bosons, Phys. Rev. A 86, 025601 (2012).
  • (24) A. K. Harter, A. Saxena, and Y. N. Joglekar, Fragile aspects of topological transition in lossy and parity-time symmetric quantum walks, Sci. Rep. 8, 12065 (2018).
  • (25) V. V. Konotop, J. Yang, and D. A. Zezyulin, Nonlinear waves in PT-symmetric systems, Rev. Mod. Phys. 88, 035002 (2016).
  • (26) D. Smirnova, D. Leykam, Y. Chong, and Y. Kivshar, Nonlinear topological photonics, Appl. Phys. Rev. 7, 021306 (2020).
  • (27) Z.-Y. Xue and Y. Hu, Topological Photonics on Superconducting Quantum Circuits with Parametric Couplings, Adv. Quant. Techn. 4, 2100017 (2021).
  • (28) P. Delplace, D. Ullmo, and G. Montambaux, Zak phase and the existence of edge states in graphene, Phys. Rev. B 84, 195452 (2011).
  • (29) J. K. Asboth, L. Oroszlany, and A. Palyi, A short course on topological insulators, Lecture Notes in Physics, Vol. 919 (Springer International Publishing, Cham, 2016).
  • (30) L.-H. Li, Z.-H. Xu, and S. Chen, Topological phases of generalized Su-Schrieffer-Heeger models, Phys. Rev. B 89, 085111 (2014).
  • (31) S. Lieu, Topological phases in the non-Hermitian Su-Schrieffer-Heeger model, Phys. Rev. B 97, 045106 (2018).
  • (32) S. Longhi, Probing one-dimensional topological phases in waveguide lattices with broken chiral symmetry, Opt. Lett. 43, 4639 (2018).
  • (33) L. Jin and Z. Song, Bulk-boundary correspondence in a non-Hermitian system in one dimension with chiral inversion symmetry, Phys. Rev. B 99, 081103(R) (2019).
  • (34) C. Yuce, Spontaneous topological pumping in non-Hermitian systems, Phys. Rev. A 99, 032109 (2019).
  • (35) L. Du, J.-H. Wu, M. Artoni, and G. C. La Rocca, Fractional quantum transport and staggered topological transition in a lossy trimerized lattice, Phys. Rev. A 100, 052102 (2019).
  • (36) L. Du, J.-H. Wu, M. Artoni, and G. C. La Rocca, Phase-dependent topological interface state and spatial adiabatic passage in a generalized Su-Schrieffer-Heeger model, Phys. Rev. A 100, 012112 (2019).
  • (37) 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).
  • (38) E. J. Bergholtz, J. C. Budich, and F. K. Kunst, Exceptional topology of non-Hermitian systems, Rev. Mod. Phys. 93, 015005 (2021).
  • (39) M. J. Ablowitz and J. F. Ladik, Nonlinear differential–difference equations and Fourier analysis, J. Math. Phys. 17, 1011 (1976).
  • (40) M. Salerno, Quantum deformations of the discrete nonlinear Schrödinger equation, Phys. Rev. A 46, 6856 (1992).
  • (41) M. Salerno, A new method to solve the quantum Ablowitz-Ladik system, Phys. Lett. A 162, 381 (1992).