]These authors contributed equally to this work. ]These authors contributed equally to this work.

Universal scaling of quantum state transport in one-dimensional topological chain under nonadiabatic dynamics

Lingzi Huang [    Menghua Deng [    Chen Sun [email protected]    Fuxiang Li [email protected] School of Physics and Electronics, Hunan University, Changsha 410082, China
Abstract

When a system is driven across a continuous phase transition, the density of topological defects demonstrates a power-law scaling behavior versus the quenching rate, as predicted by Kibble-Zurek mechanism. In this study, we generalized this idea and address the scaling of quantum state transport in a one-dimensional topological system subject to a linear drive through its topological quantum phase transition point. We illustrate the power-law dependencies of the quantum state’s transport distance, width, and peak magnitude on the driving velocity. Crucially, the power-law exponents are distinct for the edge state and bulk state. Our results offer a novel perspective on quantum state transfer and enriches the field of Kibble-Zurek behaviors and nonadiabatic quantum dynamics.

I Introduction

Quantum state transfer in quantum networks is of crucial importance in quantum control and large-scale quantum information processing Bennett1992 ; Bennett1993 . To achieve efficient quantum state transfer across quantum networks, various transfer mechanisms have been proposed, leveraging different physical systems. Among these mechanisms, most rely on one-dimensional quantum chains with either static or dynamic parameters Bose2003 ; Christandl2004 . These schemes can be realized in a variety of physical systems including photonic lattice Bellec2012 ; Perez2013 ; Chapman2016 , acoustic system Shen2019 , nitrogen-vacancy centers in diamond Yao2011 , superconducting qubit circuits Mei2018 ; Tsomokos2010 , chains of tunnel coupled quantum dots Petrosyan2006 , driven optical lattices Chen2011 , NMR Cappellaro2007 , and so on.

The efficiency of quantum state transfer schemes is predominantly dictated by two main factors: the transfer speed and the fidelity Ashhab2012 ; Caneva2009 ; Deffner2017 ; Yung2006 ; Zhang2018 . Often, these factors are at odds, presenting a trade-off between high speed and high fidelity. While adiabatic quantum evolution can ensure perfect state transfer, it typically results in slow transfer speeds. Conversely, pursuing high transfer speeds usually incurs non-adiabatic excitations, compromising the fidelity of the quantum stateGras ; Knysh ; Barends . To reconcile the two conflicting factors, the emerging field of topological states of matter provides a promising platform Hasan2010 ; Qi2011 ; Sarma2006 ; Gardas . For instance, recent studies of high-fidelity quantum state transfer have utilized the Su-Schrieffer-Heeger (SSH) model to realize robust and fast quantum state transfer protocols Estarellas ; Lang ; Longhi2019 ; Longhi20192 ; DAngelis2020 ; Xu2023 . The SSH model is particularly advantageous as it features edge states that are inherently resistant to external disorder, owing to topological protection Kitaev ; Asboth .

Most quantum state transfer protocols operate under adiabatic conditions, typically requiring an instantaneous energy gap to achieve high fidelity. A prominent example of this is Thouless pum**, which adiabatically transfers a quantized number of electrons over a periodic cycleThouless1983 . It poses an intriguing question: what occurs when the system is nonadiabatically driven across a quantum critical point where the gap closes Gras ; Knysh ; Barends . When a system is quenched across a continuous phase transition, topological excitations are formed and the density of these excitations demonstrates a universal power-law scaling relationship with the quenching speed. This phenomena, known as Kibble-Zurek mechanism (KZM) Kibble1976 ; Kibble1980 ; Zurek1985 ; Zurek1996 ; Lee2015 , has been experimentally verified in a variety of platforms Deutschlander ; Maegochi ; Du2023 ; Weiler ; Lamporesi ; Navon ; Anquez ; Ko2019 ; Yi2020 ; Keesling ; Ebadi . Recent research has expanded on the KZM, uncovering that universal characteristics are also present in rapid quench regime, in full counting statistics of defects and significant fluctuations, thereby enhancing the understanding of nonequilibrium and nonadiabatic dynamics Campo2018 ; Zeng2023 ; Balducci2023 . This raises an interesting question: are there more physical quantities that exhibit a scaling relationship with the quenching rate if the system traverses a critical point with a closing gap?

In this paper, we investigate the scaling behavior of quantum state transport in a one-dimensional topological chain when the system is linearly driven across the topological quantum phase transition point. We focus on the transport distance, the width and the peak magnitude of the quantum state, revealing that each of the quantities exhibits a power-law scaling with the driving speed. Crucially, the power-law scaling exponents display distinct values for the edge state and bulk state. We establish that our findings are applicable not only to the Hermitian SSH model, but also to other 1D topological system such as Creutz ladder model Creutz ladder ; Jafari2019 and non-Hermitian SSH model. Our research offers a novel insight for the quantum state transfer and contributes to the broader understanding of Kibble-Zurek behaviors and nonadiabatic quantum dynamics.

This paper is organized as follows. In Section II, we introduce the quenching protocol in an SSH model. In Section III, the scaling behaviors of travel distance, the width and the peak magnitude are discussed. In Section IV, we present theoretical arguments on the scaling behaviors. Sections V and VI are on results on Creutz ladder model and non-Hermitian SSH model, respectively. Finally, discussions and conclusions are presented in Section VII. Appendices A and B are on calculations on probabilities projected to each extended eigenstate for linear quenches from an edge and on discussions of fidelity of adiabatic transfer, respectively.

II SSH model and quenching protocol

We start with the 1D SSH model to study the quantum state transport during a nonadiabatic dynamics when the system is driven across the topological phase transition point. Our results are also applicable to other 1D topological system, such as Creutz ladder model and non-Hermitian SSH model, as presented in later sections. The Hamiltonian of an open-boundary SSH chain with N𝑁Nitalic_N unit cells reads

H=n=1NJ1anbn+n=1N1J2an+1bn+h.c.,formulae-sequence𝐻superscriptsubscript𝑛1𝑁subscript𝐽1superscriptsubscript𝑎𝑛subscript𝑏𝑛superscriptsubscript𝑛1𝑁1subscript𝐽2superscriptsubscript𝑎𝑛1subscript𝑏𝑛𝑐\displaystyle H=\sum_{n=1}^{N}J_{1}a_{n}^{\dagger}b_{n}+\sum_{n=1}^{N-1}J_{2}a% _{n+1}^{\dagger}b_{n}+h.c.,italic_H = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_h . italic_c . , (1)

where ansubscriptsuperscript𝑎𝑛a^{\dagger}_{n}italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (ansubscript𝑎𝑛a_{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) and bnsubscriptsuperscript𝑏𝑛b^{\dagger}_{n}italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (bnsubscript𝑏𝑛b_{n}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) are the creation (annihilation) operators on A𝐴Aitalic_A and B𝐵Bitalic_B sublattice in n𝑛nitalic_n-th unit cells. J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are intracell and intercell couplings, respectively, and h.c.formulae-sequence𝑐h.c.italic_h . italic_c . denotes Hermitian conjugate of all previous terms. It has been very well understood that the system undergoes a phase transition from topologically nontrivial phase (J1<J2subscript𝐽1subscript𝐽2J_{1}<J_{2}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) to trivial phase (J1>J2subscript𝐽1subscript𝐽2J_{1}>J_{2}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) with gap closing at J1=J2subscript𝐽1subscript𝐽2J_{1}=J_{2}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

We consider a quenching process that an initial state |ψ(ti)ket𝜓subscript𝑡𝑖|\psi(t_{i})\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ evolves under the Shrödinger equation iddt|ψ=H(t)|ψ𝑖𝑑𝑑𝑡ket𝜓𝐻𝑡ket𝜓i\frac{d}{dt}|\psi\rangle=H(t)|\psi\rangleitalic_i divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG | italic_ψ ⟩ = italic_H ( italic_t ) | italic_ψ ⟩ to a final state |ψ(tf)ket𝜓subscript𝑡𝑓|\psi(t_{f})\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ with a time-dependent Hamiltonian H(t)𝐻𝑡H(t)italic_H ( italic_t ). We consider two different quench protocols. The first one is called “linear quench” with the couplings varying with time:

J1=βt,J2=1βt.formulae-sequencesubscript𝐽1𝛽𝑡subscript𝐽21𝛽𝑡\displaystyle J_{1}=\beta t,~{}~{}J_{2}=1-\beta t.italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_β italic_t , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 - italic_β italic_t . (2)

Quench is taken from ti=0subscript𝑡𝑖0t_{i}=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 to tf=1/βsubscript𝑡𝑓1𝛽t_{f}=1/\betaitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 / italic_β, with β>0𝛽0\beta>0italic_β > 0 characterizing the speed of the quench. Such a quench connects the two fully dimerized limits J1=0subscript𝐽10J_{1}=0italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, J2=1subscript𝐽21J_{2}=1italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 and J1=1subscript𝐽11J_{1}=1italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1, J2=0subscript𝐽20J_{2}=0italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0. The other is called “periodic quench” which consists of two successive linear quenches, the quench (2) and its mirror reflection about the t=1/β𝑡1𝛽t=1/\betaitalic_t = 1 / italic_β line, such that the system returns to the original Hamiltonian after one period, and can thus be comparable to a Thouless pum** process if quenching rate β𝛽\betaitalic_β is sufficiently small.

We will consider two different initial states:

|ψ(ti)=|1,Aket𝜓subscript𝑡𝑖ket1𝐴\displaystyle|\psi(t_{i})\rangle=|1,A\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ = | 1 , italic_A ⟩ (3)

which is located at the left edge, and

|ψ(ti)=(|n,B+|n+1,A)/2ket𝜓subscript𝑡𝑖ket𝑛𝐵ket𝑛1𝐴2\displaystyle|\psi(t_{i})\rangle=(|n,B\rangle+|n+1,A\rangle)/\sqrt{2}| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ = ( | italic_n , italic_B ⟩ + | italic_n + 1 , italic_A ⟩ ) / square-root start_ARG 2 end_ARG (4)

which is located in the bulk of the chain, and is far away from the edge. For example, we can choose n=N/2𝑛𝑁2n=N/2italic_n = italic_N / 2. The two initial states are both eigenstates of the initial Hamiltonian H(ti)𝐻subscript𝑡𝑖H(t_{i})italic_H ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ).

III Scaling of travel distances

We first consider the linear quench (2). In the adiabatic limit β0𝛽0\beta\rightarrow 0italic_β → 0, since the edge state is always gapped (even though small) from other states, |1,Aket1𝐴|1,A\rangle| 1 , italic_A ⟩ evolves adiabatically to the right edge state |N,Aket𝑁𝐴|N,A\rangle| italic_N , italic_A ⟩. In the opposite diabatic limit β𝛽\beta\rightarrow\inftyitalic_β → ∞, |1,Aket1𝐴|1,A\rangle| 1 , italic_A ⟩ remains unchanged simply because there is no time to evolve. We are interested in the intermediate region between these two limits, where the final state is expected to be distributed over the chain. By numerically solving the time-dependent Schrödinger equation, we calculated |ψ(tf)ket𝜓subscript𝑡𝑓|\psi(t_{f})\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ for an open SSH chain with N=1000𝑁1000N=1000italic_N = 1000 unit cells from an initial state at the left edge (Eq. 3) or at the middle of the chain (Eq. 4). To characterize the final state, we define the final probabilities

pn,±=|n,±|ψ(tf)|2,\displaystyle p_{n,\pm}=|\langle n,\pm|\psi(t_{f})\rangle|^{2},italic_p start_POSTSUBSCRIPT italic_n , ± end_POSTSUBSCRIPT = | ⟨ italic_n , ± | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (5)

which are the probabilities of final state projected to final localized eigenstates |n,±=(|n,A±|n,B)/2ket𝑛plus-or-minusplus-or-minusket𝑛𝐴ket𝑛𝐵2|n,\pm\rangle=(|n,A\rangle\pm|n,B\rangle)/\sqrt{2}| italic_n , ± ⟩ = ( | italic_n , italic_A ⟩ ± | italic_n , italic_B ⟩ ) / square-root start_ARG 2 end_ARG. Note that at final time tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, the Hamiltonian is dimerized, and |n,±ket𝑛plus-or-minus|n,\pm\rangle| italic_n , ± ⟩ are its eigenstates with energies being ±1plus-or-minus1\pm 1± 1. Calculations show that for |ψ(ti)=|1,Aket𝜓subscript𝑡𝑖ket1𝐴|\psi(t_{i})\rangle=|1,A\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ = | 1 , italic_A ⟩ we always have pn,+=pn,subscript𝑝𝑛subscript𝑝𝑛p_{n,+}=p_{n,-}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_n , - end_POSTSUBSCRIPT, which is due to chiral symmetry of the system. For bulk state (4) with n=500𝑛500n=500italic_n = 500, pn,subscript𝑝𝑛p_{n,-}italic_p start_POSTSUBSCRIPT italic_n , - end_POSTSUBSCRIPT is much smaller than pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT due to the fact that the initial state |ψ(ti)ket𝜓subscript𝑡𝑖|\psi(t_{i})\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ is an eigenstate of the initial Hamiltonian with positive eigenvalue. Thus, it suffices to focus on pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT. One sees qualitatively different behaviors for the two types of initial states: for evolutions from the edge (Fig. 1(a)), pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT shows a smooth profile with a single peak, resembling a coherent state; whereas for evolutions from the middle (Fig. 1(b)), pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT develops oscillations whose strength increases when moving away from the initial position, until a peak emerges followed by a sharp fall-off to zero at large n𝑛nitalic_n. We also note that, as depicted in Fig. 1(c), the transport doesn’t start until near the critical point at time t=0.5/β𝑡0.5𝛽t=0.5/\betaitalic_t = 0.5 / italic_β. This observation is helpful in understanding the scaling behaviors later.

Refer to caption
Figure 1: Distribution of probabilities pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT vs. n𝑛nitalic_n for different initial states. (a) and (b) are the distributions of final probabilities for initial edge state (3) and for bulk state (4), respectively. In (b) the n𝑛nitalic_n axis is shifted such that the initial bulk state is at n=0𝑛0n=0italic_n = 0. (c) Probability profiles pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT at different times for a fixed quenching rate β=5×104𝛽5superscript104\beta=5\times 10^{-4}italic_β = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. Each profile is stretched vertically so the maximum pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT becomes 1111. (d) Probability profiles pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT of initial edge state for different values of β𝛽\betaitalic_β collapse to a single curve after scaling of n𝑛nitalic_n and pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT.
Refer to caption
Figure 2: Scalings in the open SSH chain for N=1000𝑁1000N=1000italic_N = 1000: (a) scaling of travel distance, (b) scaling of the width, (c) scaling of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT at the peak. Numerical results are represented by plot markers, blue dots for initial edge and red triangles for initial bulk state. Solid lines are power-law fittings. (d) Probability p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT to return to the edge for a periodic quench from an initial edge state. The inset shows that p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT oscillates with a period 2π/β2𝜋𝛽2\pi/\beta2 italic_π / italic_β.

The scaling behaviors of the initial edge state can be summarized in Fig. 1(d), in which, the probability distribution pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT for different values of quenching rate β𝛽\betaitalic_β collapse to a single curve after one rescales n𝑛nitalic_n and pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT by β𝛽\betaitalic_β. This indicates that all the quantities, such as travel distance and peak magnitude, that are related to the transported quantum state would exhibit power-law relations with quenching rate, as illustrated in Fig. 2. First, we consider how far the state travels under the quench. We define the position of the peak as the traveled distance d𝑑ditalic_d. In Fig. 2(a), d𝑑ditalic_d exhibits a power-law scaling for both initial edge state and bulk state dβνproportional-to𝑑superscript𝛽𝜈d\propto\beta^{-\nu}italic_d ∝ italic_β start_POSTSUPERSCRIPT - italic_ν end_POSTSUPERSCRIPT, but with different scaling exponents. For edge state, ν0.61𝜈0.61\nu\approx 0.61italic_ν ≈ 0.61, while for bulk state, ν1.03𝜈1.03\nu\approx 1.03italic_ν ≈ 1.03. For initial edge state, one can also study the width of the profile of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT, defined by the standard variance wedge=[n(nd)2pn,+]1/2subscript𝑤𝑒𝑑𝑔𝑒superscriptdelimited-[]subscript𝑛superscript𝑛𝑑2subscript𝑝𝑛12w_{edge}=[\sum_{n}(n-d)^{2}p_{n,+}]^{1/2}italic_w start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT = [ ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_n - italic_d ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. As plotted in Fig. 2(b), it also exhibits a power-law scaling with the same exponent as the travel distance: wedgeβ0.61proportional-tosubscript𝑤𝑒𝑑𝑔𝑒superscript𝛽0.61w_{edge}\propto\beta^{-0.61}italic_w start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 0.61 end_POSTSUPERSCRIPT.

Moreover, in Fig. 2(c) the maxima of probability distribution pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT of both initial edge state and initial bulk state are plotted as a function of quenching rate β𝛽\betaitalic_β. One sees that again the maximum exhibits a power-law scaling with β𝛽\betaitalic_β, but with different scaling exponents for the two types of initial states. For initial edge state, the exponent is roughly 0.610.610.610.61, while for initial bulk state, 2/3232/32 / 3 can fit the numerical results very well.

We note that in all our calculations pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT decay fast to zero for n𝑛nitalic_n sufficiently far away from the initial positions. Under periodic boundary conditions, we find that the results for evolutions from the bulk also fulfills the same scalings, dPBCβ1.03proportional-tosubscript𝑑𝑃𝐵𝐶superscript𝛽1.03d_{PBC}\propto\beta^{-1.03}italic_d start_POSTSUBSCRIPT italic_P italic_B italic_C end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 1.03 end_POSTSUPERSCRIPT and pn,+,maxβ2/3proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽23p_{n,+,max}\propto\beta^{2/3}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT. We note that different boundary conditions or different length of lattice don’t change these results as long as the end of the chain is not reached by the state evolution.

For periodic quench, after one periodic circle, one would expect that the state would transfer from one edge state to another edge state located at the other end of the chain, if the quench is slow enough so that the system is under adiabatic regime, which is called Thouless pum**. However, when crossing a topological phase transition point with gap closing, near this gap-closing point the adiabatic condition always breaks no matter how slow the quench is; this follows from the adiabatic theorem which states that conditions of an adiabatic evolution are that the quench is slow enough and that the spectrum of the system is always gapped. Thus, the adiabatic condition cannot be fully satisfied, leading to nonzero populations to excited states, and thus the quantized state transfer cannot be realized. Here in order to quantify the state transfer, we examine the final probability at the first lattice point p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT after one periodic quench. Plotting p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT vs. β𝛽\betaitalic_β for quenches from an edge in Fig. 2(d), we see that p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT oscillates between zero and an upper bound p1,A,maxsubscript𝑝1𝐴𝑚𝑎𝑥p_{1,A,max}italic_p start_POSTSUBSCRIPT 1 , italic_A , italic_m italic_a italic_x end_POSTSUBSCRIPT which scales as β0.3superscript𝛽0.3\beta^{0.3}italic_β start_POSTSUPERSCRIPT 0.3 end_POSTSUPERSCRIPT. Plotting p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT vs. 1/β1𝛽1/\beta1 / italic_β shows that this oscillation has a fixed period 2π/β2𝜋𝛽2\pi/\beta2 italic_π / italic_β (inset of Fig. 2(d)).

It’s interesting to see whether the observed scaling laws apply to other quench protocols. For evolution from an edge, we considered a quench with a sinusoidal shape,

J1=sin2(βt),J2=cos2(βt),formulae-sequencesubscript𝐽1superscript2𝛽𝑡subscript𝐽2superscript2𝛽𝑡\displaystyle J_{1}=\sin^{2}(\beta t),~{}~{}J_{2}=\cos^{2}(\beta t),italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_β italic_t ) , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_β italic_t ) , (6)

from ti=0subscript𝑡𝑖0t_{i}=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 to tf=π/(2β)subscript𝑡𝑓𝜋2𝛽t_{f}=\pi/(2\beta)italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_π / ( 2 italic_β ). We find that the behavior of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT is similar to a linear quench shown in Fig. 1.

For evolution from the bulk, we considered the above sinusoidal quench and a sudden quench where J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are both set to non-zero constants from t=0𝑡0t=0italic_t = 0 to tf=1/βsubscript𝑡𝑓1𝛽t_{f}=1/\betaitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 / italic_β, and find the same scaling laws dbulkβ1proportional-tosubscript𝑑𝑏𝑢𝑙𝑘superscript𝛽1d_{bulk}\propto\beta^{-1}italic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and pn,+,maxβ2/3proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽23p_{n,+,max}\propto\beta^{2/3}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT.

IV Theory of quenches

In this section, we present theoretical arguments on the scaling behaviors of the quantum state transport under nonadiabatic quench dynamics discussed above.

IV.1 Theory of quench from an edge

To understand the scaling behaviors, we first look at the spectrum of the system. For a chain with odd number of sites, the eigenvalues of the Hamiltonian have analytical forms Coutant_2020 : there is a single edge state with zero energy, and 2N22𝑁22N-22 italic_N - 2 extended states with energies

E=±ϵj=±|J2+J1eikj|𝐸plus-or-minussubscriptitalic-ϵ𝑗plus-or-minussubscript𝐽2subscript𝐽1superscript𝑒𝑖subscript𝑘𝑗E=\pm\epsilon_{j}=\pm|J_{2}+J_{1}e^{ik_{j}}|italic_E = ± italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ± | italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT |

where kj=πj/Nsubscript𝑘𝑗𝜋𝑗𝑁k_{j}=\pi j/Nitalic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_π italic_j / italic_N and j=1,2,,N1𝑗12𝑁1j=1,2,\ldots,N-1italic_j = 1 , 2 , … , italic_N - 1. As a result of such a spectrum, during an evolution the edge state is gapped from all the extended states except near time t=tf/2=1/(2β)𝑡subscript𝑡𝑓212𝛽t=t_{f}/2=1/(2\beta)italic_t = italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2 = 1 / ( 2 italic_β ). For an evolution starting from the left edge, we expect that during the quench the state should remain near the left edge until close to tf/2subscript𝑡𝑓2t_{f}/2italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2 when the gaps become minimal, see Fig. 1(c). At tf/2subscript𝑡𝑓2t_{f}/2italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2, a portion of the state transfers to each extended state, which then travels at the group velocity of the extended state

vk=dϵj/dk=J1J2sin(kj)/ϵj,subscript𝑣𝑘𝑑subscriptitalic-ϵ𝑗𝑑𝑘subscript𝐽1subscript𝐽2subscript𝑘𝑗subscriptitalic-ϵ𝑗\displaystyle v_{k}=d\epsilon_{j}/dk=J_{1}J_{2}\sin(k_{j})/\epsilon_{j},italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_d italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / italic_d italic_k = italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_sin ( italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) / italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (7)

and the whole state is a superposition of traveling of each extended state. In Appendix A, we further justify this picture by calculating probabilities projected to each extended eigenstate.

Based on this picture, we can estimate the travel distance dedgesubscript𝑑𝑒𝑑𝑔𝑒d_{edge}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT. Note that the Hamiltonian depends linearly on time and thus constitutes a multi-level Landau-Zener (LZ) problem, which may not be solved exactly. The traditional LZ problem is to find the transition probabilities from an initial state to each final energy state. The exactly solvable models are rare and have been found to exist only in some special forms Sinitsyn2014 ; Sinitsyn2016 ; Li2017 ; Li2018 ; LZSM . Here, instead of looking for an exact solution, we provide an Ansatz that can be fitted by numerical results. We observe that the transition probability to each final state (denoted as pksubscript𝑝𝑘p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT note-pk ) is solely determined by a dimensionless parameter Δk2/βsuperscriptsubscriptΔ𝑘2𝛽\Delta_{k}^{2}/\betaroman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_β, where Δk=ϵk=cos(k/2)subscriptΔ𝑘subscriptitalic-ϵ𝑘𝑘2\Delta_{k}=\epsilon_{k}=\cos(k/2)roman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = roman_cos ( italic_k / 2 ) is the gap between an extended state and the edge state at t=tf/2𝑡subscript𝑡𝑓2t=t_{f}/2italic_t = italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2. One can use the following Ansatz:

pk=c3(1ec1Δk2β)ec2Δk2β.subscript𝑝𝑘subscript𝑐31superscript𝑒subscript𝑐1superscriptsubscriptΔ𝑘2𝛽superscript𝑒subscript𝑐2superscriptsubscriptΔ𝑘2𝛽\displaystyle p_{k}=c_{3}\left(1-e^{-\frac{c_{1}\Delta_{k}^{2}}{\beta}}\right)% e^{-\frac{c_{2}\Delta_{k}^{2}}{\beta}}.italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT . (8)

c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c2subscript𝑐2c_{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are two fitting parameters, and c3subscript𝑐3c_{3}italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is an overall normalization factor. Comparison with numerical results are presented in Appendix A. Note that pksubscript𝑝𝑘p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is small both at small and at large ΔksubscriptΔ𝑘\Delta_{k}roman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, and it reaches a peak at Δk2/β1superscriptsubscriptΔ𝑘2𝛽1\Delta_{k}^{2}/\beta\approx 1roman_Δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_β ≈ 1. The wavevector corresponding to this peak is k=2arccosβ𝑘2𝛽k=2\arccos\sqrt{\beta}italic_k = 2 roman_arccos square-root start_ARG italic_β end_ARG, which can be used to estimate the travel distance as

dedge=1/(2β)1/β|vk|𝑑tsubscript𝑑𝑒𝑑𝑔𝑒superscriptsubscript12𝛽1𝛽subscript𝑣𝑘differential-d𝑡\displaystyle d_{edge}=\int_{1/(2\beta)}^{1/\beta}|v_{k}|dtitalic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT 1 / ( 2 italic_β ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT | italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_d italic_t (9)

assuming that the traveling roughly starts at time 1/(2β)12𝛽1/(2\beta)1 / ( 2 italic_β ) and ends at time 1/β1𝛽1/\beta1 / italic_β. This integral can be analytically performed, giving

dedge=18β(2β1βarccosh1β11β).subscript𝑑𝑒𝑑𝑔𝑒18𝛽2𝛽1𝛽arccosh1𝛽11𝛽\displaystyle d_{edge}=\frac{1}{8\sqrt{\beta}}\left(\frac{2-\beta}{1-\beta}% \operatorname{arccosh}\frac{1}{\sqrt{\beta}}-\sqrt{\frac{1}{1-\beta}}\right).italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 8 square-root start_ARG italic_β end_ARG end_ARG ( divide start_ARG 2 - italic_β end_ARG start_ARG 1 - italic_β end_ARG roman_arccosh divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_β end_ARG end_ARG - square-root start_ARG divide start_ARG 1 end_ARG start_ARG 1 - italic_β end_ARG end_ARG ) . (10)

The green curve in Fig. 2(a) is Eq. (10), which agrees well with the exact dedgesubscript𝑑𝑒𝑑𝑔𝑒d_{edge}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT, and at large β𝛽\betaitalic_β it correctly predicts the drop of dedgesubscript𝑑𝑒𝑑𝑔𝑒d_{edge}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT below the power-law scaling.

The power-law scaling of dedgesubscript𝑑𝑒𝑑𝑔𝑒d_{edge}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT can now be understood. Since β1much-less-than𝛽1\beta\ll 1italic_β ≪ 1 in almost the whole range of β𝛽\betaitalic_β considered, kee** the leading order in β𝛽\betaitalic_β gives

dedge18β(log4β1).subscript𝑑𝑒𝑑𝑔𝑒18𝛽4𝛽1\displaystyle d_{edge}\approx\frac{1}{8\sqrt{\beta}}\left(\log\frac{4}{\beta}-% 1\right).italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ≈ divide start_ARG 1 end_ARG start_ARG 8 square-root start_ARG italic_β end_ARG end_ARG ( roman_log divide start_ARG 4 end_ARG start_ARG italic_β end_ARG - 1 ) . (11)

Thus, dedgesubscript𝑑𝑒𝑑𝑔𝑒d_{edge}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT scales as dedgeβ1/2log(1/β)proportional-tosubscript𝑑𝑒𝑑𝑔𝑒superscript𝛽121𝛽d_{edge}\propto\beta^{-1/2}\log(1/\beta)italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT roman_log ( 1 / italic_β ), which is visually close to a power law dedgeβνproportional-tosubscript𝑑𝑒𝑑𝑔𝑒superscript𝛽𝜈d_{edge}\propto\beta^{-\nu}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - italic_ν end_POSTSUPERSCRIPT with some ν>1/2𝜈12\nu>1/2italic_ν > 1 / 2 within a range of β𝛽\betaitalic_β.

IV.2 Theory for quench from the bulk

As mentioned before, for evolution from the bulk we could consider a chain with periodic boundaries. Now the Hamiltonian can be reduced to a 2×2222\times 22 × 2 matrix in the momentum space labeled by k=2πj/N𝑘2𝜋𝑗𝑁k=2\pi j/Nitalic_k = 2 italic_π italic_j / italic_N (j=0,1,,N1𝑗01𝑁1j=0,1,\ldots,N-1italic_j = 0 , 1 , … , italic_N - 1), with eigenenergies E=±ϵk=±|J1+J2eik|𝐸plus-or-minussubscriptitalic-ϵ𝑘plus-or-minussubscript𝐽1subscript𝐽2superscript𝑒𝑖𝑘E=\pm\epsilon_{k}=\pm|J_{1}+J_{2}e^{-ik}|italic_E = ± italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ± | italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k end_POSTSUPERSCRIPT |. Let’s denote the corresponding eigenvectors in the k𝑘kitalic_kth-block as |ψk,±ketsubscript𝜓𝑘plus-or-minus|\psi_{k,\pm}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , ± end_POSTSUBSCRIPT ⟩. The initial localized state |ψ(ti)=(|ni,B+|ni+1,A)/2ket𝜓subscript𝑡𝑖ketsubscript𝑛𝑖𝐵ketsubscript𝑛𝑖1𝐴2|\psi(t_{i})\rangle=(|n_{i},B\rangle+|n_{i}+1,A\rangle)/\sqrt{2}| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ = ( | italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_B ⟩ + | italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + 1 , italic_A ⟩ ) / square-root start_ARG 2 end_ARG can be written as a superposition of all eigenstates with positive energies:

|ψ(ti)=(1/N)keikni|k|ψk,+.ket𝜓subscript𝑡𝑖1𝑁subscript𝑘tensor-productsuperscript𝑒𝑖𝑘subscript𝑛𝑖ket𝑘ketsubscript𝜓𝑘\displaystyle|\psi(t_{i})\rangle=(1/\sqrt{N})\sum_{k}e^{-ikn_{i}}|k\rangle% \otimes|\psi_{k,+}\rangle.| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ = ( 1 / square-root start_ARG italic_N end_ARG ) ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | italic_k ⟩ ⊗ | italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩ . (12)

During the quench each |k|ψk,+tensor-productket𝑘ketsubscript𝜓𝑘|k\rangle\otimes|\psi_{k,+}\rangle| italic_k ⟩ ⊗ | italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩ travels with velocity vk=ϵj/ksubscript𝑣𝑘subscriptitalic-ϵ𝑗𝑘v_{k}=\partial\epsilon_{j}/\partial kitalic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∂ italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / ∂ italic_k, and accumulates to a distance 01/βvk𝑑tsuperscriptsubscript01𝛽subscript𝑣𝑘differential-d𝑡\int_{0}^{1/\beta}v_{k}dt∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_d italic_t at the final time note-LZ . Note that unlike evolution from the edge, the traveling starts immediately at t=0𝑡0t=0italic_t = 0. We can estimate the travel distance dbulksubscript𝑑𝑏𝑢𝑙𝑘d_{bulk}italic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT as the largest distance traveled by all the eigenstates, which simply gives rise to dbulk=c/βsubscript𝑑𝑏𝑢𝑙𝑘𝑐𝛽d_{bulk}=c/\betaitalic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT = italic_c / italic_β obtained by a change of variable tt/β𝑡𝑡𝛽t\rightarrow t/\betaitalic_t → italic_t / italic_β. Here, c𝑐citalic_c is a constant and numerically determined to be around 0.240.240.240.24.

For the understanding of the magnitude of peak pn,+,maxsubscript𝑝𝑛𝑚𝑎𝑥p_{n,+,max}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT, we consider the corresponding amplitude that can be approximated by assuming adiabatic evolution note-adiabatic :

nmax,+|ψ(tf)=(1/N)keikniei(knmaxφk),\displaystyle\langle n_{max},+|\psi(t_{f})\rangle=(1/N)\sum_{k}e^{-ikn_{i}}e^{% i(kn_{max}-\varphi_{k})},⟨ italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT , + | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ = ( 1 / italic_N ) ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( italic_k italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT - italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT , (13)

where φk=01/βϵk𝑑tsubscript𝜑𝑘superscriptsubscript01𝛽subscriptitalic-ϵ𝑘differential-d𝑡\varphi_{k}=\int_{0}^{1/\beta}\epsilon_{k}dtitalic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_d italic_t is the adiabatic phase accumulated during the evolution of |ψk,+ketsubscript𝜓𝑘|\psi_{k,+}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩. The magnitude can be estimated by the stationary phase approximation. Let’s consider Taylor expansion of φksubscript𝜑𝑘\varphi_{k}italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT near kmaxsubscript𝑘𝑚𝑎𝑥k_{max}italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT, which is the momentum labeling each final state with maximum transition probability. Therefore, one has (2φk/k2)kmax=0subscriptsuperscript2subscript𝜑𝑘superscript𝑘2subscript𝑘𝑚𝑎𝑥0(\partial^{2}\varphi_{k}/\partial k^{2})_{k_{max}}=0( ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / ∂ italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0. Since φk/k=01/βvktsubscript𝜑𝑘𝑘superscriptsubscript01𝛽subscript𝑣𝑘𝑡\partial\varphi_{k}/\partial k=\int_{0}^{1/\beta}v_{k}t∂ italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / ∂ italic_k = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_t, we have (φk/k)kmax=dbulk=nmaxnisubscriptsubscript𝜑𝑘𝑘subscript𝑘𝑚𝑎𝑥subscript𝑑𝑏𝑢𝑙𝑘subscript𝑛𝑚𝑎𝑥subscript𝑛𝑖(\partial\varphi_{k}/\partial k)_{k_{max}}=d_{bulk}=n_{max}-n_{i}( ∂ italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / ∂ italic_k ) start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Thus, the exponent in nmax,+|ψ(tf)\langle n_{max},+|\psi(t_{f})\rangle⟨ italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT , + | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ is dominated by the third order expansion (kkmax)3superscript𝑘subscript𝑘𝑚𝑎𝑥3(k-k_{max})^{3}( italic_k - italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT term. It’s easy to see that φk1/βproportional-tosubscript𝜑𝑘1𝛽\varphi_{k}\propto 1/\betaitalic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∝ 1 / italic_β, so (3φk/k3)k=kmax=C/βsubscriptsuperscript3subscript𝜑𝑘superscript𝑘3𝑘subscript𝑘𝑚𝑎𝑥𝐶𝛽(\partial^{3}\varphi_{k}/\partial k^{3})_{k=k_{max}}=C/\beta( ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / ∂ italic_k start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_k = italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_C / italic_β, where C𝐶Citalic_C is a constant. Replacing the summation ksubscript𝑘\sum_{k}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT by integration 𝑑k/(2π/N)differential-d𝑘2𝜋𝑁\int dk/(2\pi/N)∫ italic_d italic_k / ( 2 italic_π / italic_N ), we arrive at

nmax,+|ψ(tf)12πdke(iC6β)(kkmax)3,\displaystyle\langle n_{max},+|\psi(t_{f})\approx\frac{1}{2\pi}\int_{-\infty}^% {\infty}dke^{-(i\frac{C}{6\beta})(k-k_{max})^{3}},⟨ italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT , + | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ≈ divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_k italic_e start_POSTSUPERSCRIPT - ( italic_i divide start_ARG italic_C end_ARG start_ARG 6 italic_β end_ARG ) ( italic_k - italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , (14)

which scales as β1/3proportional-toabsentsuperscript𝛽13\propto\beta^{1/3}∝ italic_β start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT, and thus pn,+,max=|nmax,+|ψ(tf)|2β2/3p_{n,+,max}=|\langle n_{max},+|\psi(t_{f})\rangle|^{2}\propto\beta^{2/3}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT = | ⟨ italic_n start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT , + | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∝ italic_β start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT.

IV.3 Theory of periodic quench

We now consider the probability p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT to return to the left edge in a periodic quench, which can be considered as consisting of two linear quenches. The return probability is given by p1,A=|k(Ak,++Ak,)|2subscript𝑝1𝐴superscriptsubscript𝑘subscript𝐴𝑘subscript𝐴𝑘2p_{1,A}=|\sum_{k}(A_{k,+}+A_{k,-})|^{2}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT = | ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_A start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT + italic_A start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where Ak,±subscript𝐴𝑘plus-or-minusA_{k,\pm}italic_A start_POSTSUBSCRIPT italic_k , ± end_POSTSUBSCRIPT is the amplitude contributed from an extended state labelled by k𝑘kitalic_k and ±plus-or-minus\pm±. It can be written as Ak,±=pkpkeiφk,±subscript𝐴𝑘plus-or-minussubscript𝑝𝑘superscriptsubscript𝑝𝑘superscript𝑒𝑖subscript𝜑𝑘plus-or-minusA_{k,\pm}=\sqrt{p_{k}}\sqrt{p_{k}^{\prime}}e^{-i\varphi_{k,\pm}}italic_A start_POSTSUBSCRIPT italic_k , ± end_POSTSUBSCRIPT = square-root start_ARG italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG square-root start_ARG italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_i italic_φ start_POSTSUBSCRIPT italic_k , ± end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, where pksuperscriptsubscript𝑝𝑘p_{k}^{\prime}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is the transition probability from the extended state to the edge state during the second linear quench, and φk,±subscript𝜑𝑘plus-or-minus\varphi_{k,\pm}italic_φ start_POSTSUBSCRIPT italic_k , ± end_POSTSUBSCRIPT is the phase for this amplitude. The sum in p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT can be simplified by noticing that the multistate LZ model corresponding to the considered linear quench belongs to bipartite models studied in cross-2017 . According to cross-2017 , the transition probability matrix of such a model is symmetric, so pk=pksuperscriptsubscript𝑝𝑘subscript𝑝𝑘p_{k}^{\prime}=p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. Moreover, chiral symmetry further implies φk,+=φk,subscript𝜑𝑘subscript𝜑𝑘\varphi_{k,+}=-\varphi_{k,-}italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT = - italic_φ start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT. Thus, we have

p1,A=(2kpkcosφk,+)2.subscript𝑝1𝐴superscript2subscript𝑘subscript𝑝𝑘subscript𝜑𝑘2\displaystyle p_{1,A}=\left(2\sum_{k}p_{k}\cos\varphi_{k,+}\right)^{2}.italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT = ( 2 ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_cos italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (15)

We may approximate φk,+subscript𝜑𝑘\varphi_{k,+}italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT at kπsimilar-to𝑘𝜋k\sim\piitalic_k ∼ italic_π by the adiabatic phase accumulated during the quench, namely, the area under the curve of ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT in an E𝐸Eitalic_E vs.t𝑡titalic_t diagram. Since the peak of pksubscript𝑝𝑘p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is at k=2arccosβ𝑘2𝛽k=2\arccos\sqrt{\beta}italic_k = 2 roman_arccos square-root start_ARG italic_β end_ARG, which for β1much-less-than𝛽1\beta\ll 1italic_β ≪ 1 is close to π𝜋\piitalic_π, we only need to calculate φk,+subscript𝜑𝑘\varphi_{k,+}italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT for kπsimilar-to𝑘𝜋k\sim\piitalic_k ∼ italic_π, which is simply φk,+1/(2β)3/(2β)ϵk𝑑t=1/(2β)subscript𝜑𝑘superscriptsubscript12𝛽32𝛽subscriptitalic-ϵ𝑘differential-d𝑡12𝛽\varphi_{k,+}\approx\int_{1/(2\beta)}^{3/(2\beta)}\epsilon_{k}dt=1/(2\beta)italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ≈ ∫ start_POSTSUBSCRIPT 1 / ( 2 italic_β ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / ( 2 italic_β ) end_POSTSUPERSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_d italic_t = 1 / ( 2 italic_β ). Therefore, we arrive at the observed result that p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT oscillates with a period 2π/β2𝜋𝛽2\pi/\beta2 italic_π / italic_β. In fact, what takes place here is the phenomenon of Landau-Zener-Stückelberg interferometry LZS-interf-2010 : an interference between the total amplitude of all positive-energy extended states kAk,+subscript𝑘subscript𝐴𝑘\sum_{k}A_{k,+}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT and that of all negative-energy extended states kAk,subscript𝑘subscript𝐴𝑘\sum_{k}A_{k,-}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT leads to a periodic dependence of the return probability on 1/β1𝛽1/\beta1 / italic_β. A maximum is reached whenever kAk,+subscript𝑘subscript𝐴𝑘\sum_{k}A_{k,+}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT and kAk,subscript𝑘subscript𝐴𝑘\sum_{k}A_{k,-}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT are in constructive interference, namely, when arg(kAk,+)=0subscript𝑘subscript𝐴𝑘0\arg(\sum_{k}A_{k,+})=0roman_arg ( ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ) = 0. Thus, at a given β𝛽\betaitalic_β the upper bound of p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT reads p1,A,max=|2kpkeiφk,+|2subscript𝑝1𝐴𝑚𝑎𝑥superscript2subscript𝑘subscript𝑝𝑘superscript𝑒𝑖subscript𝜑𝑘2p_{1,A,max}=\left|2\sum_{k}p_{k}e^{i\varphi_{k,+}}\right|^{2}italic_p start_POSTSUBSCRIPT 1 , italic_A , italic_m italic_a italic_x end_POSTSUBSCRIPT = | 2 ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. A calculation of p1,A,maxsubscript𝑝1𝐴𝑚𝑎𝑥p_{1,A,max}italic_p start_POSTSUBSCRIPT 1 , italic_A , italic_m italic_a italic_x end_POSTSUBSCRIPT using Eq. (15) with numerically obtained pksubscript𝑝𝑘p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and φk,+subscript𝜑𝑘\varphi_{k,+}italic_φ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT agrees well with the upper bound of p1,Asubscript𝑝1𝐴p_{1,A}italic_p start_POSTSUBSCRIPT 1 , italic_A end_POSTSUBSCRIPT in Fig. 2(d).

V Scaling behaviors in Creutz ladder model

Refer to caption
Figure 3: (a) The diagram depicts the structure of the Creutz ladder model with A and B the two sublattices. W𝑊Witalic_W, K𝐾Kitalic_K, and Ke±iθ𝐾superscript𝑒plus-or-minus𝑖𝜃Ke^{\pm i\theta}italic_K italic_e start_POSTSUPERSCRIPT ± italic_i italic_θ end_POSTSUPERSCRIPT are the vertical, horizontal and, diagonal hop** integrals. (b) The spectrum of E𝐸Eitalic_E vs. βt𝛽𝑡\beta titalic_β italic_t for quenching M𝑀Mitalic_M and K𝐾Kitalic_K of the Creutz ladder model. (c) The spectrum of E𝐸Eitalic_E vs. βt𝛽𝑡\beta titalic_β italic_t for quenching the complex phase θ𝜃\thetaitalic_θ of the Creutz ladder model.

We have considered the SSH model as an example to study the scaling behaviors of quantum state transfer under nonadiabatic quench dynamics. To illustrate that the scaling behavior is not restricted to some special model but is universal, we now consider another model originally proposed by Creutz in Creutz ladder . The Creutz model is also a 1D chain consisting of two lattice points A𝐴Aitalic_A and B𝐵Bitalic_B, as depicted in Fig. 3. The Hamiltonian reads

HCreutz=i=1N[K(anbn+1+bnan+1)\displaystyle H_{Creutz}=-\sum_{i=1}^{N}[K(a^{\dagger}_{n}b_{n+1}+b_{n}^{% \dagger}a_{n+1})italic_H start_POSTSUBSCRIPT italic_C italic_r italic_e italic_u italic_t italic_z end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT [ italic_K ( italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT + italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT )
+K(eiθanan+1+eiθbnbn+1)𝐾superscript𝑒𝑖𝜃subscriptsuperscript𝑎𝑛subscript𝑎𝑛1superscript𝑒𝑖𝜃superscriptsubscript𝑏𝑛subscript𝑏𝑛1\displaystyle\qquad+K(e^{-i\theta}a^{\dagger}_{n}a_{n+1}+e^{i\theta}b_{n}^{% \dagger}b_{n+1})+ italic_K ( italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT + italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT )
+Manbn+h.c.],\displaystyle\qquad+Ma_{n}^{\dagger}b_{n}+h.c.],+ italic_M italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_h . italic_c . ] , (16)

where n𝑛nitalic_n label the unit cells with a total of number of N𝑁Nitalic_N, A𝐴Aitalic_A and B𝐵Bitalic_B label the sublattice, M𝑀Mitalic_M, K𝐾Kitalic_K, Keiθ𝐾superscript𝑒𝑖𝜃Ke^{-i\theta}italic_K italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT are the vertical, diagonal, and horizontal hop** integrals, respectively. The phase factor e±iθsuperscript𝑒plus-or-minus𝑖𝜃e^{\pm i\theta}italic_e start_POSTSUPERSCRIPT ± italic_i italic_θ end_POSTSUPERSCRIPT mimics the presence of a magnetic field which pierces the ladder and supplies a magnetic flux θ/π𝜃𝜋\theta/\piitalic_θ / italic_π per plaquette. ansubscriptsuperscript𝑎𝑛a^{\dagger}_{n}italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (ansubscript𝑎𝑛a_{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) and bnsubscriptsuperscript𝑏𝑛b^{\dagger}_{n}italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (bnsubscript𝑏𝑛b_{n}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) are the creation (annihilation) operators on the sublattice A𝐴Aitalic_A and B𝐵Bitalic_B in the n𝑛nitalic_n-th unit cell, respectively. We consider two different quenching protocols. The first one is quenching M𝑀Mitalic_M and K𝐾Kitalic_K

M=βt,K=1βt,formulae-sequence𝑀𝛽𝑡𝐾1𝛽𝑡\displaystyle M=\beta t,~{}~{}K=1-\beta t,italic_M = italic_β italic_t , italic_K = 1 - italic_β italic_t , (17)

with fixing θ=π/2𝜃𝜋2\theta=-\pi/2italic_θ = - italic_π / 2 and time changing from ti=0subscript𝑡𝑖0t_{i}=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 to tf=1/βsubscript𝑡𝑓1𝛽t_{f}=1/\betaitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 / italic_β. The second one is quenching θ𝜃\thetaitalic_θ in the following way

θ=βtπ/2(π2,π2),𝜃𝛽𝑡𝜋2𝜋2𝜋2\displaystyle\theta=\beta t-\pi/2\in(-\frac{\pi}{2},\frac{\pi}{2}),italic_θ = italic_β italic_t - italic_π / 2 ∈ ( - divide start_ARG italic_π end_ARG start_ARG 2 end_ARG , divide start_ARG italic_π end_ARG start_ARG 2 end_ARG ) , (18)

with fixing M=0𝑀0M=0italic_M = 0, K=1𝐾1K=1italic_K = 1 and time changing from ti=0subscript𝑡𝑖0t_{i}=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 to tf=π/βsubscript𝑡𝑓𝜋𝛽t_{f}=\pi/\betaitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_π / italic_β. Both of the two quench protocols are described by the quenching rate β𝛽\betaitalic_β. The instantaneous eigenvalues of the Hamiltonian as functions of time t𝑡titalic_t are plotted in Fig. 3.

We consider two different initial localized states. One is a plaquette-blocked state

|ψ(ti)n=12(ian+bn+an+1ibn+1)|0subscriptket𝜓subscript𝑡𝑖𝑛12𝑖superscriptsubscript𝑎𝑛superscriptsubscript𝑏𝑛superscriptsubscript𝑎𝑛1𝑖superscriptsubscript𝑏𝑛1ket0\displaystyle|\psi(t_{i})\rangle_{n}=\frac{1}{2}(-ia_{n}^{\dagger}+b_{n}^{% \dagger}+a_{n+1}^{\dagger}-ib_{n+1}^{\dagger})|0\rangle| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( - italic_i italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + italic_a start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - italic_i italic_b start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) | 0 ⟩ (19)

localized in the middle of the chain with n=N/2𝑛𝑁2n=N/2italic_n = italic_N / 2. The other is a topological edge state located at the left edge

|L=12(a1ib1)|0.ket𝐿12superscriptsubscript𝑎1𝑖superscriptsubscript𝑏1ket0\displaystyle|L\rangle=\frac{1}{\sqrt{2}}(a_{1}^{\dagger}-ib_{1}^{\dagger})|0\rangle.| italic_L ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - italic_i italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) | 0 ⟩ . (20)

Note that these initial states are eigenstates of the initial Hamiltonian.

The two quenching protocols start with the same initial parameters but go through different phases. Under open boundary conditions, the initial parameters are M=0𝑀0M=0italic_M = 0, K=1𝐾1K=1italic_K = 1 and θ=π/2𝜃𝜋2\theta=-\pi/2italic_θ = - italic_π / 2, there are two localized chiral zero-energy modes at the left edge and the right edge. For the first quench protocol with quenching M𝑀Mitalic_M and K𝐾Kitalic_K, the system undergoes a phase transition at time t=0.65/β𝑡0.65𝛽t=0.65/\betaitalic_t = 0.65 / italic_β, after which the system enters into a topologically trivial phase with no edge states(as shown in Fig. 3(b)). For the second quench protocol with quenching only θ𝜃\thetaitalic_θ, one can see that there are always two zero-energy modes as shown in Fig. 3(c), even though the system crosses a quantum critical point at θc=0subscript𝜃𝑐0\theta_{c}=0italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0 with gap closing.

We now discuss the scaling behavior of quantum state transfer under quench dynamics of the Creutz ladder model by numerically solving the time-dependent Schrödinger equation for a chain with N=1000𝑁1000N=1000italic_N = 1000. First we study the first kind of quench protocol that quenches M𝑀Mitalic_M and K𝐾Kitalic_K. We plot the probability profile pnsubscript𝑝𝑛p_{n}italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of the final state in Fig. 4(a) for the two different initial states, the left edge state |Lket𝐿|L\rangle| italic_L ⟩ and the plaquette-block state |ψ(ti)n,+subscriptket𝜓subscript𝑡𝑖𝑛|\psi(t_{i})\rangle_{n,+}| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT localized in the middle of the chain with n=500𝑛500n=500italic_n = 500. Similar to the case of the SSH model, we can see that the distribution of final states exhibit quite different profiles for the two different initial states. For initial edge state, the final state has a smooth profile for different values of β𝛽\betaitalic_β, while for initial plaquette-block state, the shape of final state is not smooth, but shows two sharp peaks.

Refer to caption
Figure 4: Profile of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT and scalings of quenching M𝑀Mitalic_M and K𝐾Kitalic_K in the open Creutz ladder for N=1000𝑁1000N=1000italic_N = 1000 for different initial states: (a)-(c) left edge state |Lket𝐿|L\rangle| italic_L ⟩ (d)-(f) bulk platform |ψ(ti)nsubscriptket𝜓subscript𝑡𝑖𝑛|\psi(t_{i})\rangle_{n}| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. (a), (d) are distribution diagrams for two cases at different β𝛽\betaitalic_β distinguished by black, blue and red dots. Scalings of travel distances and the maximum of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT for |Lket𝐿|L\rangle| italic_L ⟩ provided in (b) and (c) shows dedgeβ0.63proportional-tosubscript𝑑𝑒𝑑𝑔𝑒superscript𝛽0.63d_{edge}\propto\beta^{-0.63}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 0.63 end_POSTSUPERSCRIPT and pn,+,maxβ0.6proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽0.6p_{n,+,max}\propto\beta^{0.6}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT 0.6 end_POSTSUPERSCRIPT. (e) and (h) show that bulk platform state transmits at the scaling as dβ1.08proportional-to𝑑superscript𝛽1.08d\propto\beta^{-1.08}italic_d ∝ italic_β start_POSTSUPERSCRIPT - 1.08 end_POSTSUPERSCRIPT and pn,+,maxβ23proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽23p_{n,+,max}\propto\beta^{\frac{2}{3}}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT divide start_ARG 2 end_ARG start_ARG 3 end_ARG end_POSTSUPERSCRIPT.

Further, we study the scaling of travel distances d𝑑ditalic_d and the maximum of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT as functions of quenching rate β𝛽\betaitalic_β. Comparing with the result in SSH model, for quenching from edge state, the travel distance and the maximum of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT obey almost the same power-law scaling that dedgeβ0.63proportional-tosubscript𝑑𝑒𝑑𝑔𝑒superscript𝛽0.63d_{edge}\propto\beta^{-0.63}italic_d start_POSTSUBSCRIPT italic_e italic_d italic_g italic_e end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 0.63 end_POSTSUPERSCRIPT and pn,+,maxβ0.6proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽0.6p_{n,+,max}\propto\beta^{0.6}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT 0.6 end_POSTSUPERSCRIPT. Additionally, the scalings for evolution from |ψ(ti)nsubscriptket𝜓subscript𝑡𝑖𝑛|\psi(t_{i})\rangle_{n}| italic_ψ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⟩ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT are almost in accordance with that in the case of evolution from SSH middle state, which present dbulkplatformβ1.08proportional-tosubscript𝑑𝑏𝑢𝑙𝑘𝑝𝑙𝑎𝑡𝑓𝑜𝑟𝑚superscript𝛽1.08d_{bulkplatform}\propto\beta^{-1.08}italic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k italic_p italic_l italic_a italic_t italic_f italic_o italic_r italic_m end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT - 1.08 end_POSTSUPERSCRIPT and pn,+,maxβ32proportional-tosubscript𝑝𝑛𝑚𝑎𝑥superscript𝛽32p_{n,+,max}\propto\beta^{\frac{3}{2}}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT ∝ italic_β start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT.

For the second quench protocol, unfortunately, if the initial state is on the edge, e.g. |Lket𝐿|L\rangle| italic_L ⟩, the state doesn’t move and is always localized on the original location. For the initial plaquette-block state, one obtains the same scaling exponents for the distance and peak of transported wavefunction.

VI scaling behavior for non-Hermitian SSH model

We consider a non-Hermitian SSH model with imaginary on-site energy iγ𝑖𝛾i\gammaitalic_i italic_γ and iγ𝑖𝛾-i\gamma- italic_i italic_γ:

H=𝐻absent\displaystyle H=italic_H = n=1N[(J1anbn+J2an+1bn+h.c.)\displaystyle\sum_{n=1}^{N}[(J_{1}a_{n}^{\dagger}b_{n}+J_{2}a_{n+1}^{\dagger}b% _{n}+h.c.)∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT [ ( italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_h . italic_c . ) (21)
+iγananiγbnbn].\displaystyle+i\gamma a_{n}^{\dagger}a_{n}-i\gamma b_{n}^{\dagger}b_{n}].+ italic_i italic_γ italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_i italic_γ italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ] .

Here, we still quench J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in such a protocol:

J1=βt,J2=1βt,formulae-sequencesubscript𝐽1𝛽𝑡subscript𝐽21𝛽𝑡\displaystyle J_{1}=\beta t,~{}~{}J_{2}=1-\beta t,italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_β italic_t , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 - italic_β italic_t , (22)

from time ti=0subscript𝑡𝑖0t_{i}=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 to tf=1/βsubscript𝑡𝑓1𝛽t_{f}=1/\betaitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 / italic_β, with β𝛽\betaitalic_β the quenching rate. Even with the presence of imaginary on-site energy, at initial time, the system is dimerized, and with two edge states located at the two ends of the chain. At final time tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, the system is also dimerized, with eigenenergies: ε±=±1γ2subscript𝜀plus-or-minusplus-or-minus1superscript𝛾2\varepsilon_{\pm}=\pm\sqrt{1-\gamma^{2}}italic_ε start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT = ± square-root start_ARG 1 - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, and eigenstates

|±=12[|n,A+(±1γ2iγ)|n,B].ketplus-or-minus12delimited-[]ket𝑛𝐴plus-or-minus1superscript𝛾2𝑖𝛾ket𝑛𝐵\displaystyle|\pm\rangle=\frac{1}{\sqrt{2}}[|n,A\rangle+(\pm\sqrt{1-\gamma^{2}% }-i\gamma)|n,B\rangle].| ± ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ | italic_n , italic_A ⟩ + ( ± square-root start_ARG 1 - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - italic_i italic_γ ) | italic_n , italic_B ⟩ ] .

During the quench, the system undergoes two exceptional points at J1J2=±γsubscript𝐽1subscript𝐽2plus-or-minus𝛾J_{1}-J_{2}=\pm\gammaitalic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ± italic_γ. The instantaneous eigenenergies with open boundary condition are plotted in Fig. 5(a) and (b). The probability distributions pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT for the initial edge state after linear quench dynamics are numerically calculated for different values of quench rate β𝛽\betaitalic_β. The profile is similar to the Hermitian case, but with a very large magnitude of the order of eγ/βsuperscript𝑒𝛾𝛽e^{\gamma/\beta}italic_e start_POSTSUPERSCRIPT italic_γ / italic_β end_POSTSUPERSCRIPT due to the non-Hermitian term. Nevertheless, after rescaling the probability with a factor eγ/βsuperscript𝑒𝛾𝛽e^{-\gamma/\beta}italic_e start_POSTSUPERSCRIPT - italic_γ / italic_β end_POSTSUPERSCRIPT, one can still obtain a very well defined probability profile as shown in Fig. 5(c) for different values of β𝛽\betaitalic_β. The same as in the Hermitian SSH model and Creutz ladder model, we still have scaling behaviors for the travel distance d𝑑ditalic_d and peak magnitude pn,+,maxsubscript𝑝𝑛𝑚𝑎𝑥p_{n,+,max}italic_p start_POSTSUBSCRIPT italic_n , + , italic_m italic_a italic_x end_POSTSUBSCRIPT with the latter rescaled by eγ/βsuperscript𝑒𝛾𝛽e^{-\gamma/\beta}italic_e start_POSTSUPERSCRIPT - italic_γ / italic_β end_POSTSUPERSCRIPT. The scaling exponent is the same 0.610.610.610.61, which indicates a universal value.

Refer to caption
Figure 5: Energy spectrum of a non-Hermitian SSH model and the scaling behavior of travelled quantum state. (a)-(b) The real part and imaginary part of instantaneous energy levels as functions of time. (c) Rescaled probability profile for different values of quench rate β𝛽\betaitalic_β. (d) Scalings of travel distances and the maximum of pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT for initial edge state. Numbers of unit cells are taken as N=50𝑁50N=50italic_N = 50 for (a) and (b), and N=500𝑁500N=500italic_N = 500 for (c) and (d).

VII Discussions and Conclusions

We have considered the dynamical behaviors of quantum state transport in a 1D topological chain which is driven across the topological phase transition with gap closing. We find that the state transport exhibits universal power-law scaling behaviors with quenching rate, which enriches the Kibble-Zurek phenomena. More importantly, starting from edge state or bulk state will produce distinct scaling exponents. After a periodic quench circle, Thouless pum** does not hold if the topological transition is crossed, but rather the returning probability to the initial edge state is nonzero and exhibits also a power-law scaling with quenching rate. Our results are of broad interest in nonequilibrium quantum statistical mechanics, connecting quantum state transfer with the breakdown of adiabatic dynamics, and should find broad applications in quantum information, quantum annealing, ultracold atom physics, and the study of critical phenomena.

Acknowledgements.
This paper was supported by the National Key Research and Development Program of the Ministry of Science and Technology (Grant No. 2021YFA1200700), the National Natural Science Foundation of China (Grants No. 12275075, and No. 12105094), and the Fundamental Research Funds for the Central Universities of China.

Appendix A: Probabilities projected to extended eigenstates for linear quenches from an edge

In the main text, for linear quenches from the edge, our theoretical estimate of the travel distance is based on the picture that the edge state transfers to each extended state near t=tf/2𝑡subscript𝑡𝑓2t=t_{f}/2italic_t = italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2, which then travels at its group velocity. Here we justify this picture by calculating probabilities projected to each extended eigenstate.

We work on a chain with an odd number of sites, whose eigenvalues and eigenstates have analytical forms. The eigenvalues from lowest to highest (assuming that J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT have the same sign) are Coutant_2020 :

ϵj=|J1+J2eikj|,for j=1,2,,N1,formulae-sequencesubscriptitalic-ϵ𝑗subscript𝐽1subscript𝐽2superscript𝑒𝑖subscript𝑘𝑗for 𝑗12𝑁1\displaystyle\epsilon_{j}=-|J_{1}+J_{2}e^{-ik_{j}}|,\quad\textrm{for }j=1,2,% \ldots,N-1,italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = - | italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | , for italic_j = 1 , 2 , … , italic_N - 1 ,
ϵN=0,subscriptitalic-ϵ𝑁0\displaystyle\epsilon_{N}=0,italic_ϵ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = 0 ,
ϵj=|J1+J2eikj|,for j=N+1,,2N1,formulae-sequencesubscriptitalic-ϵ𝑗subscript𝐽1subscript𝐽2superscript𝑒𝑖subscript𝑘𝑗for 𝑗𝑁12𝑁1\displaystyle\epsilon_{j}=|J_{1}+J_{2}e^{-ik_{j}}|,\quad\textrm{for }j=N+1,% \ldots,2N-1,italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = | italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | , for italic_j = italic_N + 1 , … , 2 italic_N - 1 , (A1)

where kj=πj/Nsubscript𝑘𝑗𝜋𝑗𝑁k_{j}=\pi j/Nitalic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_π italic_j / italic_N (Note that, unlike in the main text, we used the integer j𝑗jitalic_j to label the states, which now to range from 1111 to 2N12𝑁12N-12 italic_N - 1. The definition of ϵjsubscriptitalic-ϵ𝑗\epsilon_{j}italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is also different from ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT; it can now take negative values.) The corresponding normalized eigenstates are:

|ψj=1N[n=1Nsin(nkj+ϕj)|n,A\displaystyle|\psi_{j}\rangle=\frac{1}{\sqrt{N}}\left[-\sum_{n=1}^{N}\sin(nk_{% j}+\phi_{j})|n,A\rangle\right.| italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG [ - ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_sin ( italic_n italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | italic_n , italic_A ⟩
+n=1N1sin(nkj)|n,B],for j<N,\displaystyle\left.+\sum_{n=1}^{N-1}\sin(nk_{j})|n,B\rangle\right],\quad% \textrm{for }j<N,+ ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT roman_sin ( italic_n italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | italic_n , italic_B ⟩ ] , for italic_j < italic_N , (A2)
|ψN=1(J1/J2)21(J1/J2)2Nn=1N(J1J2)n1|n,A,ketsubscript𝜓𝑁1superscriptsubscript𝐽1subscript𝐽221superscriptsubscript𝐽1subscript𝐽22𝑁superscriptsubscript𝑛1𝑁superscriptsubscript𝐽1subscript𝐽2𝑛1ket𝑛𝐴\displaystyle|\psi_{N}\rangle=\sqrt{\frac{1-(J_{1}/J_{2})^{2}}{1-(J_{1}/J_{2})% ^{2N}}}\sum_{n=1}^{N}\left(-\frac{J_{1}}{J_{2}}\right)^{n-1}|n,A\rangle,| italic_ψ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ⟩ = square-root start_ARG divide start_ARG 1 - ( italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 1 - ( italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( - divide start_ARG italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT | italic_n , italic_A ⟩ , (A3)
|ψj=1N[n=1Nsin(nkj+ϕj)|n,A\displaystyle|\psi_{j}\rangle=\frac{1}{\sqrt{N}}\left[\sum_{n=1}^{N}\sin(nk_{j% }+\phi_{j})|n,A\rangle\right.| italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG [ ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_sin ( italic_n italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | italic_n , italic_A ⟩
+n=1N1sin(nkj)|n,B],for j>N,\displaystyle\left.+\sum_{n=1}^{N-1}\sin(nk_{j})|n,B\rangle\right],\quad% \textrm{for }j>N,+ ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT roman_sin ( italic_n italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | italic_n , italic_B ⟩ ] , for italic_j > italic_N , (A4)

where ϕj=arg(J1+J2eikj)subscriptitalic-ϕ𝑗subscript𝐽1subscript𝐽2superscript𝑒𝑖subscript𝑘𝑗\phi_{j}=\arg(J_{1}+J_{2}e^{-ik_{j}})italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = roman_arg ( italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) is a phase shift on sublattice A𝐴Aitalic_A. |ψNketsubscript𝜓𝑁|\psi_{N}\rangle| italic_ψ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ⟩ is the edge state, which is localized at the left (right) end of the chain if J1<J2subscript𝐽1subscript𝐽2J_{1}<J_{2}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (J1>J2subscript𝐽1subscript𝐽2J_{1}>J_{2}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT). All other eigenstates are extended; note that they depend on time through ϕjsubscriptitalic-ϕ𝑗\phi_{j}italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

With these analytical expressions, one can readily calculate |ψj|ψ(t)|2superscriptinner-productsubscript𝜓𝑗𝜓𝑡2|\langle\psi_{j}|\psi(t)\rangle|^{2}| ⟨ italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_ψ ( italic_t ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, i.e. probabilities of the evolving state projected to each extended eigenstate at any time. We looked at |ψj|ψ(t)|2superscriptinner-productsubscript𝜓𝑗𝜓𝑡2|\langle\psi_{j}|\psi(t)\rangle|^{2}| ⟨ italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_ψ ( italic_t ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the final time tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT of a linear quench, which we denote as pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. We observe that pj=p2Njsubscript𝑝𝑗subscript𝑝2𝑁𝑗p_{j}=p_{2N-j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT 2 italic_N - italic_j end_POSTSUBSCRIPT, which originates from chiral symmetry, so it suffices to consider only the j>N𝑗𝑁j>Nitalic_j > italic_N states. Fig. A1 shows pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT vs. ΔjsubscriptΔ𝑗\Delta_{j}roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (Δj=|cos(kj/2)|subscriptΔ𝑗subscript𝑘𝑗2\Delta_{j}=|\cos(k_{j}/2)|roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = | roman_cos ( italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / 2 ) | is the gap between the j𝑗jitalic_jth extended state and the edge state at t=tf/2𝑡subscript𝑡𝑓2t=t_{f}/2italic_t = italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2) at β=5×104𝛽5superscript104\beta=5\times 10^{-4}italic_β = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. We see that as ΔjsubscriptΔ𝑗\Delta_{j}roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT increases, pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT increases from zero, reaches a peak and then decreases to zero. We find that for different β𝛽\betaitalic_β the profiles of pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT vs. ΔjsubscriptΔ𝑗\Delta_{j}roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT can be fitted by a function of the form

pj=c3(1ec1Δj2β)ec2Δj2β.subscript𝑝𝑗subscript𝑐31superscript𝑒subscript𝑐1superscriptsubscriptΔ𝑗2𝛽superscript𝑒subscript𝑐2superscriptsubscriptΔ𝑗2𝛽\displaystyle p_{j}=c_{3}\left(1-e^{-\frac{c_{1}\Delta_{j}^{2}}{\beta}}\right)% e^{-\frac{c_{2}\Delta_{j}^{2}}{\beta}}.italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( 1 - italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT . (A5)

The fitting parameters c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c2subscript𝑐2c_{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT depend slightly on β𝛽\betaitalic_β, and c3subscript𝑐3c_{3}italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is an overall normalization factor. At β=5×104𝛽5superscript104\beta=5\times 10^{-4}italic_β = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT reaches its maximum at Δj2/β=0.967superscriptsubscriptΔ𝑗2𝛽0.967\Delta_{j}^{2}/\beta=0.967roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_β = 0.967.

Refer to caption
Figure A1: pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT vs. ΔjsubscriptΔ𝑗\Delta_{j}roman_Δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT at β=5×104𝛽5superscript104\beta=5\times 10^{-4}italic_β = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT for N=1000𝑁1000N=1000italic_N = 1000 sites for evolutions from an edge. The blue dots are exact results, and the red line is a fitting by Eq. (A5) with c1=1.03subscript𝑐11.03c_{1}=1.03italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1.03, c2=0.657subscript𝑐20.657c_{2}=0.657italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.657 and c3=0.0856subscript𝑐30.0856c_{3}=0.0856italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0.0856.
Refer to caption
Figure A2: pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT vs. n𝑛nitalic_n at β=5×104𝛽5superscript104\beta=5\times 10^{-4}italic_β = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT for N=1000𝑁1000N=1000italic_N = 1000 sites for evolutions from an edge. The blue dots are exact results, and the red line is from the calculation using pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT as described in the text, with each wave package taken as a rectangular function (pj/40)rect[(n1dj)/40]subscript𝑝𝑗40rect𝑛1subscript𝑑𝑗40(p_{j}/40)\operatorname{rect}[(n-1-d_{j})/40]( italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / 40 ) roman_rect [ ( italic_n - 1 - italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) / 40 ].

We further use pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT to calculate the final state’s profile pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT, by assuming that the j𝑗jitalic_jth eigenstate travels at a speed |vk|=|ϵj/k|subscript𝑣𝑘subscriptitalic-ϵ𝑗𝑘|v_{k}|=|\partial\epsilon_{j}/\partial k|| italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | = | ∂ italic_ϵ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / ∂ italic_k | and finally produces a wave package with a strength pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT centered at dj=1/(2β)1/β|vk|𝑑tsubscript𝑑𝑗superscriptsubscript12𝛽1𝛽subscript𝑣𝑘differential-d𝑡d_{j}=\int_{1/(2\beta)}^{1/\beta}|v_{k}|dtitalic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT 1 / ( 2 italic_β ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT | italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_d italic_t, and the final profile pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT is a summation of all these wave packages. Each wave package has a certain shape with a certain width which our theory does not predict, and we simply take it to be a rectangular function (pj/W)rect[(n1dj)/W]subscript𝑝𝑗𝑊rect𝑛1subscript𝑑𝑗𝑊(p_{j}/W)\operatorname{rect}[(n-1-d_{j})/W]( italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / italic_W ) roman_rect [ ( italic_n - 1 - italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) / italic_W ], where the width W𝑊Witalic_W (assumed to be the same for all j𝑗jitalic_j) is treated as a fitting parameter. Such a calculation gives a profile which reproduces well the exact pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT, as shown in Fig. A2. This further confirms the picture that the edge state transfers to each extended state and then travels at its group velocity.

Appendix B: Fidelity of adiabatic transfer

The power-law scalings we found take place when the state does not evolve to the other edge (the edges) of the chain. Namely, we have been effectively considering a semi-infinite (infinite) chain. For a finite chain with an odd number of sites, in the adiabatic limit β0𝛽0\beta\rightarrow 0italic_β → 0 we expect that |1,Aket1𝐴|1,A\rangle| 1 , italic_A ⟩ evolves adiabatically to |N,Aket𝑁𝐴|N,A\rangle| italic_N , italic_A ⟩, since the edge state is always gapped from other states. For a larger β𝛽\betaitalic_β this adiabatic transfer is not perfect, and its efficiency can be characterized by the fidelity F=|N,B|ψ(tf)|2𝐹superscriptinner-product𝑁𝐵𝜓subscript𝑡𝑓2F=|\langle N,B|\psi(t_{f})\rangle|^{2}italic_F = | ⟨ italic_N , italic_B | italic_ψ ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In Fig. B1 we plot F𝐹Fitalic_F vs. β𝛽\betaitalic_β for a chain with N=100𝑁100N=100italic_N = 100 with an odd number of sites. We found that F𝐹Fitalic_F can be fitted by a simple formula:

F=[max(12e2N2β,0)]2,𝐹superscriptdelimited-[]12superscript𝑒2superscript𝑁2𝛽02\displaystyle F=\left[\max\left(1-2e^{-\frac{2}{N^{2}\beta}},0\right)\right]^{% 2},italic_F = [ roman_max ( 1 - 2 italic_e start_POSTSUPERSCRIPT - divide start_ARG 2 end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_β end_ARG end_POSTSUPERSCRIPT , 0 ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (B1)

which works for different values of N𝑁Nitalic_N. The appearance of 12e2/(N2β)12superscript𝑒2superscript𝑁2𝛽1-2e^{-2/(N^{2}\beta)}1 - 2 italic_e start_POSTSUPERSCRIPT - 2 / ( italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_β ) end_POSTSUPERSCRIPT in (B1) reminds us the transition probability of staying on the middle level in the bow-tie model bow-tie , and implies that the considered SSH model with linear time-dependence may be somehow similar to the bow-tie model.

Refer to caption
Figure B1: Fidelity of adiabatic transfer in the SSH model, F𝐹Fitalic_F vs. β𝛽\betaitalic_β, for N=100𝑁100N=100italic_N = 100 for evolutions from an edge. The blue dots are exact results, and the red dashed line is Eq. (B1).

References

  • (1) C. H. Bennett and S. J. Wiesner, Communication via one- and two-particle operators on Einstein-Podolsky-Rosen states, Phys. Rev. Lett. 69, 2881 (1992).
  • (2) C. H. Bennett, G. Brassard, C. Crépeau, R. Jozsa, A. Peres, and W. K. Wootters, Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels, Phys. Rev. Lett. 70, 1895 (1993).
  • (3) S. Bose, Quantum communication through an unmodulated spin chain, Phys. Rev. Lett. 91, 207901 (2003).
  • (4) M. Christandl, N. Datta, A. Ekert, and A. J. Landahl, Perfect State Transfer in Quantum Spin Networks, Phys. Rev. Lett. 92, 187902 (2004).
  • (5) M. Bellec, G. M. Nikolopoulos, and S. Tzortzakis, Faithful communication Hamiltonian in photonic lattices, Opt. Lett. 37, 4504 (2012).
  • (6) A. Perez-Leija, R. Keil, A. Kay, H. MoyaCessa, S. Nolte, L. C. Kwek, B. M. Rodríguez-Lara, A. Szameit, amd D. N. Christodoulides, Coherent quantum transport in photonic lattices, Phys. Rev. A. 87, 012309 (2013).
  • (7) R. J. Chapman, M. Santandrea, Z. Huang, G. Corrielli, A. Crespi, M. H. Yung, R. Osellame, and A. Peruzzo, Experimental perfect state transfer of an entangled photonic qubit, Nat. Commun. 7, 11339 (2016)
  • (8) Y. X. Shen, Y. G. Peng, D. G. Zhao, X. C. Chen, J. Zhu, and X. F. Zhu, One-Way Localized Adiabatic Passage in an Acoustic System, Phys. Rev. Lett. 122, 094501 (2019).
  • (9) N. Y. Yao, L. Jiang, A. V. Gorshkov, Z.-X. Gong, A. Zhai, L. M. Duan, and M. D. Lukin, Robust Quantum State Transfer in Random Unpolarized Spin Chains, Phys. Rev. Lett. 106, 040505 (2011).
  • (10) F. Mei, G. Chen, L. Tian, S. L. Zhu, and S. Jia, Robust quantum state transfer via topological edge states in superconducting qubit chains, Phys. Rev. A. 98, 012331 (2018).
  • (11) D. I. Tsomokos, S. Ashhab, and F. Nori, Using superconducting qubit circuits to engineer exotic lattice systems, Phys. Rev. A. 82, 052311 (2010).
  • (12) D. Petrosyan and P. Lambropoulos, Coherent population transfer in a chain of tunnel coupled quantum dots, Opt. Commun. 264, 419 (2006).
  • (13) Y. A. Chen, S. Nascimbène, M. Aidelsburger, M. Atala, S. Trotzky, and I. Bloch, Controlling Correlated Tunneling and Superexchange Interactions with ac-Driven Optical Lattices, Phys. Rev. Lett. 107, 210405 (2011).
  • (14) P. Cappellaro, C. Ramanathan, and D. G. Cory, Simulations of Information Transport in Spin Chains, Phys. Rev. Lett. 99, 250506 (2007)
  • (15) S. Ashhab, P. C. de Groot, and F. Nori, Speed limits for quantum gates in multiqubit systems, Phys. Rev. A. 85, 052327 (2012).
  • (16) T. Caneva, M. Murphy, T. Calarco, R. Fazio, S. Montangero, V. Giovannetti, and G. E. Santoro, Optimal Control at the Quan- tum Speed Limit, Phys. Rev. Lett. 103, 240501 (2009).
  • (17) S. Deffner and S. Campbell, Quantum speed limits: from Heisenberg’s uncertainty principle to optimal quantum control, J. Phys. A: Math. Theor. 50, 453001 (2017).
  • (18) M. H. Yung, Quantum speed limit for perfect state transfer in one dimension, Phys. Rev. A. 74, 030303(R) (2006).
  • (19) X. M. Zhang, Z. W. Cui, X. Wang, and M. H. Yung, Automatic spin-chain learning to explore the quantum speed limit, Phys. Rev. A. 97, 052333 (2018).
  • (20) T. Graß, D. Raventós, B. Juliá-Díaz, G. Christian, and M. Lewenstein, Quantum annealing for the number-partitioning problem using a tunable spin glass of ions, Nat. Commun. 7, 11524 (2016).
  • (21) S. Knysh, Zero-temperature quantum annealing bottlenecks in the spin-glass phase, Nat. Commun. 7, 12370 (2016).
  • (22) R. Barends, A. Shabani, L. Lamata, J. Kelly, A. Mezzacapo, U. Las Heras, R. Babbush, A. G. Fowler, B. Campbell, Yu Chen, Z. Chen, B. Chiaro, A. Dunsworth, E. Jeffrey, E. Lucero, A. Megrant, J. Y. Mutus, M. Neeley, C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, D. Sank, A. Vainsencher, and John M. Martinis, Digitized adiabatic quantum computing with a superconducting circuit, Nature (London) 534, 222 (2016).
  • (23) M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).
  • (24) X. L. Qi and S. C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • (25) S. D. Sarma, M. Freedman, and C. Nayak, Topological quantum computation, Phys. Today 59(7), 32 (2006).
  • (26) B. Gardas, J. Dziarmaga, W. H. Zurek, and M. Zwolak, Defects in Quantum Computers, Sci. Rep. 8, 4539 (2018).
  • (27) M. P. Estarellas, I. D’Amico, and T. P. Spiller, Topologically protected localized states in spin chains, Sci. Rep. 7, 1 (2017).
  • (28) N. Lang and H. P. Büchler, Topological networks for quantum communication between distant qubits, npj Quantum Inf. 3, 1 (2017).
  • (29) S. Longhi, Topological pum** of edge states via adiabatic passage, Phys. Rev. B. 99, 155150 (2019).
  • (30) S. Longhi, G. L. Giorgi, and R. Zambrini, Landau zener topological quantum state transfer, Adv. Quantum Technol. 2, 1800090 (2019).
  • (31) F. M. D’Angelis, F. A. Pinheiro, D. Guéry-Odelin, S Longhi, and F. Impens, Fast and robust quantum state transfer in a topological Su-Schrieffer-Heeger chain with next-to-nearest-neighbor interactions, Phys. Rev. Research 2, 033475 (2020).
  • (32) J. Xu, F. Mei, and Y.-Q. Zhu, A universal shortcut method for state transfer in quantum spin systems, arXiv:2312.08920 [quant-ph], 14 Dec 2023.
  • (33) A. Y. Kitaev, Unpaired majorana fermions in quantum wires, Phys. Usp. 44, 131 (2001).
  • (34) J. K. Asbóth, L. Oroszlány, and A. Pályi, A Short Course on Topological Insulators, Lecture Notes in Physics Vol. 919 (2016).
  • (35) D. J. Thouless, Quantization of particle transport, Phys. Rev. B. 27, 6083 (1983).
  • (36) T. W. B. Kibble, Topology of cosmic domains and strings, J. Phys. A. 9, 1387 (1976).
  • (37) T. W. B. Kibble, Some implications of a cosmological phase transition, Phys. Rep. 67, 183 (1980).
  • (38) W. H. Zurek, Cosmological experiments in superfluid helium?, Nature (London) 317, 505 (1985).
  • (39) W. H. Zurek, Cosmological experiments in condensed matter systems, Phys. Rep. 276, 177 (1996).
  • (40) M. Lee, S. Han, and M.-S. Choi, Kibble-Zurek mechanism in a topological phase transition, Phys. Rev. B 92, 035117 (2015).
  • (41) S. Deutschländer, P. Dillmann, G. Maret, and P. Keim, Kibble-Zurek mechanism in colloidal monolayers, Proc. Natl. Acad. Sci. U.S.A. 112, 6925 (2015).
  • (42) S. Maegochi, K. Ienaga, and S. Okuma, Kibble-Zurek Mechanism for Dynamical Ordering in a Driven Vortex System, Phys. Rev. Lett. 129, 227001 (2022).
  • (43) K. Du, X. Fang, C. Won, C. De, F. T. Huang, W. Xu, H. You, F. J. Gómez-Ruiz, A. del Campo, and S. W. Cheong, Kibble-Zurek mechanism of Ising domains, Nat. Phys. 19, 1495 (2023).
  • (44) C. N. Weiler, T. W. Neely, D. R. Scherer, A. S. Bradley, M. J. Davis, and B. P. Anderson, Spontaneous vortices in the formation of Bose-Einstein condensates, Nature (London) 455, 948 (2008).
  • (45) G. Lamporesi, S. Donadello, S. Serafini, F. Dalfovo, and G. Ferrari, Spontaneous creation of Kibble-Zurek solitons in a Bose-Einstein condensate, Nat. Phys. 9, 656 (2013).
  • (46) N. Navon, A. L. Gaunt, R. P. Smith, and Z. Hadzibabic, Critical dynamics of spontaneous symmetry breaking in a homogeneous Bose gas, Science 347, 167 (2015).
  • (47) M. Anquez, B. A. Robbins, H. M. Bharath, M. Boguslawski, T. M. Hoang, and M. S. Chapman, Quantum Kibble-Zurek Mechanism in a Spin-1 Bose-Einstein Condensate, Phys. Rev. Lett. 116, 155301 (2016).
  • (48) B. Ko, J. W. Park, and Y. Shin, Kibble-Zurek universality in a strongly interacting Fermi superfluid, Nat. Phys. 15, 1227 (2019).
  • (49) C. R. Yi, S. Liu, R. H. Jiao, J. Y. Zhang, Y. S. Zhang, and S. Chen, Exploring Inhomogeneous Kibble-Zurek Mechanism in a Spin-Orbit Coupled Bose-Einstein Condensate, Phys. Rev. Lett. 125, 260603 (2020)
  • (50) A. Keesling, A. Omran, H. Levine, H. Bernien, H. Pichler, S. Choi, R. Samajdar, S. Schwartz, P. Silvi, S. Sachdev, P. Zoller, M. Endres, M. Greiner, V. Vuletić, and M. D. Lukin, Quantum Kibble-Zurek mechanism and critical dynamics on a programmable Rydberg simulator, Nature (London) 568, 207 (2019).
  • (51) S. Ebadi, T. T. Wang, H. Levine, A. Keesling, G. Semeghini, A. Omran, D. Bluvstein, R. Samajdar, H. Pichler, W. W. Ho, S. Choi, S. Sachdev, M. Greiner, V. Vuletić, and M. D. Lukin, Quantum phases of matter on a 256-atom programmable quantum simulator, Nature (London) 595, 227 (2021).
  • (52) A. del Campo, Universal Statistics of Topological Defects Formed in a Quantum Phase Transition, Phys. Rev. Lett. 121, 200601 (2018).
  • (53) HuaBi. Z, ChuanYin. X, and Adolfo del Campo, Universal Breakdown of Kibble-Zurek Scaling in Fast Quenches across a Phase Transition, Phys. Rev. Lett. 130, 060402 (2023).
  • (54) Federico, B. Mathieu Beau, Y. **g, Andrea Gambassi, and Adolfo del Campo, Large Deviations beyond the Kibble-Zurek Mechanism, Phys. Rev. Lett. 131 (23), 230401 (2023).
  • (55) M. Creutz, End states, ladder compounds, and domain-wall fermions, Phys. Rev. Lett. 83, 2636 (1999).
  • (56) R. Jafari, H. Johannesson, A. Langari, and M. A. Martin-Delgado, Quench dynamics and zero-energy modes: The case of the Creutz model, Phys. Rev. B 99, 054302 (2019).
  • (57) A. Coutant, V. Achilleos, O. Richoux, G. Theocharis, and V. Pagneux, Robustness of topological corner modes against disorder with application to acoustic networks, Phys. Rev. B. 102, 214204 (2020).
  • (58) N. A. Sinitsyn, Exact results for models of multichannel quantum nonadiabatic transitions, Phys. Rev. A. 90, 062509 (2014).
  • (59) N. A. Sinitsyn and F. Li, Solvable multistate model of Landau-Zener transitions in cavity QED, Phys. Rev. A. 93, 063859 (2016).
  • (60) F. Li, C. Sun, V. Y. Chernyak, and N. A. Sinitsyn, Multistate Landau-Zener models with all levels crossing at one point, Phys. Rev. A. 96, 022107 (2017).
  • (61) F. Li, V. Y. Chernyak, and N. A. Sinitsyn, Quantum Annealing and Thermalization: Insights from Integrability, Phys. Rev. Lett. 121, 190601 (2018).
  • (62) L. Landau, Zur Theorie der Energieubertragung II, Phys. Z. Sowj. 2, 46 (1932); C. Zener. Non-Adiabatic Crossing of Energy Levels, Proc. R. Soc. 137, 696 (1932); E. Majorana, Atomi orientati in campo magnetico variabile, Nuovo Cimento 9, 43 (1932); E. C. G. Stückelberg. Theorie der unelastischen Stösse zwischen Atomen, Helv. Phys. Acta. 5, 370 (1932).
  • (63) We found that transition probabilities from the edge state to the two extended states at the same k𝑘kitalic_k with positive and negative energies are the same, which is due to chiral symmetry. They are thus denoted by the same symbol pksubscript𝑝𝑘p_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT.
  • (64) F. Li, C. Sun, V. Y. Chernyak, and N. A. Sinitsyn, Multistate Landau-Zener models with all levels crossing at one point, Phys. Rev. A. 96, 022107 (2017).
  • (65) S. N. Shevchenko, S. Ashhab, and Franco Nori, Landau-Zener-Stuckelberg interferometry, Physics Reports, 492, 1 (2010).
  • (66) One may think that a transition within a 2×2222\times 22 × 2 bulk, namely from |ψk,+ketsubscript𝜓𝑘|\psi_{k,+}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩ to |ψk,ketsubscript𝜓𝑘|\psi_{k,-}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT ⟩ can take place, which may influence the estimate of dbulksubscript𝑑𝑏𝑢𝑙𝑘d_{bulk}italic_d start_POSTSUBSCRIPT italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT. Actually at the wavevector kmaxsubscript𝑘𝑚𝑎𝑥k_{max}italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT (which value to be determined shortly after) we will use to calculate dT,bulksubscript𝑑𝑇𝑏𝑢𝑙𝑘d_{T,bulk}italic_d start_POSTSUBSCRIPT italic_T , italic_b italic_u italic_l italic_k end_POSTSUBSCRIPT, such a transition can be ignored, due to the following reason. Each 2×2222\times 22 × 2 block can be described by a two-state LZ model H=2βsin(k/2)tσz+cos(k/2)σx𝐻2𝛽𝑘2𝑡subscript𝜎𝑧𝑘2subscript𝜎𝑥H=2\beta\sin(k/2)t\sigma_{z}+\cos(k/2)\sigma_{x}italic_H = 2 italic_β roman_sin ( italic_k / 2 ) italic_t italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + roman_cos ( italic_k / 2 ) italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, and the LZ model corresponding to the value of kmaxsubscript𝑘𝑚𝑎𝑥k_{max}italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT is in the adiabatic region since exp(2πΓ)=exp[2πcos2(kmax/2)/(4βsin(kmax/2))]12𝜋Γ2𝜋superscript2subscript𝑘𝑚𝑎𝑥24𝛽subscript𝑘𝑚𝑎𝑥2much-less-than1\exp(-2\pi\Gamma)=\exp[-2\pi\cos^{2}(k_{max}/2)/(4\beta\sin(k_{max}/2))]\ll 1roman_exp ( - 2 italic_π roman_Γ ) = roman_exp [ - 2 italic_π roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT / 2 ) / ( 4 italic_β roman_sin ( italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT / 2 ) ) ] ≪ 1 in the whole range of β𝛽\betaitalic_β considered, so transition amplitudes between |ψk,+ketsubscript𝜓𝑘|\psi_{k,+}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩ and |ψk,ketsubscript𝜓𝑘|\psi_{k,-}\rangle| italic_ψ start_POSTSUBSCRIPT italic_k , - end_POSTSUBSCRIPT ⟩ are vanishingly small.
  • (67) As argued in note-LZ , when considering pn,+subscript𝑝𝑛p_{n,+}italic_p start_POSTSUBSCRIPT italic_n , + end_POSTSUBSCRIPT near the peak, evolutions of all |k|ψk,+tensor-productket𝑘ketsubscript𝜓𝑘|k\rangle\otimes|\psi_{k,+}\rangle| italic_k ⟩ ⊗ | italic_ψ start_POSTSUBSCRIPT italic_k , + end_POSTSUBSCRIPT ⟩ states involved can be treated as adiabatic.
  • (68) V. N. Ostrovsky and H. Nakamura, Exact analytical solution of the N𝑁Nitalic_N-level Landau-Zener-type bow-tie model, J. Phys. A: Math. Gen. 30, 6939 (1997).