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

Superconducting state generated dynamically from distant pair source and drain

E. S. Ma    Z. Song songtc@nankai.edu.cn School of Physics, Nankai University, Tianjin 300071, China
Abstract

It has been well established that the origin of pp-wave superconductivity is the balance between pair creation and annihilation, described by the spin-less fermionic Kitaev model. In this work, we study the dynamics of a composite system where the pair source and drain are spatially separated by a long distance. We show that this non-Hermitian system possesses a high-order exceptional point (EP) when only a source or drain is considered. The EP dynamics provide a clear picture: A pair source can fully fill the system with pairs, while a drain can completely empty the system. When the two coexist simultaneously, the dynamics depend on the distance and the relative phase between the pair creation and annihilation terms. Analytical analysis and numerical simulation results show that the superconducting state can be dynamically established at the resonant pair source and drain: from an initial empty state to a stationary state with the maximal pair order parameter. It provides an alternative way of understanding the mechanism of the nonequilibrium superconducting state.

I Introduction

Motivated by recent advances in experimental capability kinoshita2006quantum ; trotzky2012probing ; gring2012relaxation ; schreiber2015observation ; smith2016many ; kaufman2016quantum , the nonequilibrium dynamics of quantum many-body systems has emerged as a fundamental and attractive topic in condensed-matter physics. As one of the potential applications, nonequilibrium many-body dynamics provide an alternative way to access a new exotic quantum state with energy far from the ground state Choi ; Else ; Khemani ; Lindner ; Kaneko ; Tindall ; YXMPRA ; zhang2022steady . Unlike traditional protocols based on the cooling mechanism, quench dynamics have a wide range of potential applications since they provide many ways to take a system out of equilibrium, such as applying a driving field or pumping energy or particles in the system through external reservoirs QD1 ; QD2 ; QD3 . This makes it possible to design interacting many-body systems to prepare some desirable many-body quantum states by a quenching process. Much effort kitamura2016eta ; kaneko2019photoinduced ; tindall2019heating ; fujiuchi2019photoinduced ; peronaci2020enhancement ; fujiuchi2020superconductivity ; ejima2020photoinduced ; li2020eta ; kaneko2020charge ; zhang2021eta ; zhang2020dynamical ; yang2022dynamic ; zhang2022steady has been devoted to developing various nonequilibrium protocols for the generation of the η\eta-pairing-like state yang1989eta in the Hubbard model.

In parallel, the Kitaev model is a lattice model of a pp-wave superconducting wire, which realize Majorana zero modes at the ends of the chain Kitaev . This has been demonstrated by unpaired Majorana modes exponentially localized at the ends of open Kitaev chains Sarma ; Stern ; Alicea . The main feature of this model originates from the pairing term, which violates the conservation of the fermion number but preserves its parity, leading to the superconducting phase. The amplitudes for pair creation and annihilation play an important role in the existence of the gapped superconducting phase. Compared to the Hubbard model, the Kitaev model has an advantage for the task since it is exactly solvable. In recent work on a simple 1D Kitaev model shi2022dynamic ; YXMPRA , it has been shown that a nonequilibrium superconducting state can be obtained through time evolution from an initially prepared vacuum state, providing an alternative approach to dynamically generate a superconducting state from an easily prepared trivial state.

Due to the translational symmetry of such a system, every pair term contributes equally to the formation of the superconducting state. A question arises as to what happens if there is only a single pair term in the Hamiltonian. In this work, we study the dynamics of a composite system where the pair source and drain are considered individually or spatially separated by a long distance. In the framework of quantum mechanics, it corresponds to a non-Hermitian Hamiltonian Bender ; Bender1 . Many contributions have been devoted to non-Hermitian Kitaev models Law ; Tong ; Yuce ; You ; Klett ; Menke and Ising models LCPRB ; LC2014conventional ; LC2015finite ; LC2016chern ; ZXZ within the pseudo-Hermitian framework. We show that this non-Hermitian system possesses a high-order exceptional point (EP) when only a source or drain is considered. It admits peculiar dynamics: the final state is a particular eigenstate, coalescing state Berry2004 ; Heiss2012 ; Miri2019 ; Zhang2020 . The EP dynamics provide a clear physical picture: A pair source can eventially fully fill the system with pairs, while a drain can completely empty the system. However, both final states are trivial. When the two coexist simultaneously, the dynamics depend on the distance and the relative phase between the pair creation and annihilation terms. Analytical analysis and numerical simulation results for a finite system show that a perfect superconducting state can be dynamically established at the resonant pair source and drain, i.e., an initial empty state evolves to a stationary state, which is a perfect superconducting state with the maximal pair order parameter. We consider a composite system with nonhomogenous pair terms in the present work in comparison to previous works. This provides an alternative mechanism for forming nonequilibrium superconducting state.

This paper is organized as follows. In Section II, we describe the model Hamiltonian and introduce the order parameter. In Section III, based on the exact solutions of a toy model, we demonstrate that its ground state has the maximal order parameter. In Section IV, we study the dynamics in the non-Hermitian system, where only pair creation or annihilation terms are considered. In Section V, we investigate the dynamics driven by a single source or drain. In Section VI, we investigate the dynamics driven by spatially separated source and drain at resonance. Finally, we give a summary and discussion in Section VII.

Refer to caption
Figure 1: Schematic of the 1D Kitaev models for spinless fermions with pair terms across two adjacent sites, red and green dimers indicating specially separated pair creation (source) and annihilation (drain), respectively. Two resonant pair terms are embedded in a (a) ring lattice and (b) open chain. The goal of this work is to investigate the effect of the setup on the nonequilibrium state after sufficiently long time.

II Hamiltonian and order parameter

We start with a generalized Hamiltonian

H\displaystyle H =\displaystyle= HT+HP,\displaystyle H_{\mathrm{T}}+H_{\mathrm{P}}, (1)
HT\displaystyle H_{\mathrm{T}} =\displaystyle= iTj=1Ncjcj+1+H.c.+μj=1Nnj,\displaystyle-iT\sum\limits_{j=1}^{N}c_{j}^{{\dagger}}c_{j+1}+\mathrm{H.c.}+\mu\sum\limits_{j=1}^{N}n_{j}, (2)
HP\displaystyle H_{\mathrm{P}} =\displaystyle= j=1,r>0N(fjrcjcj+r+gjrcj+rcj),\displaystyle\sum\limits_{j=1,r>0}^{N}\left(f_{j}^{r}c_{j}^{{\dagger}}c_{j+r}^{{\dagger}}+g_{j}^{r}c_{j+r}c_{j}\right), (3)

where the distribution functions fjrf_{j}^{r} and gjrg_{j}^{r} determine the location of creation and annihilation of a pair with size rr, respectively. In the case with fjr=gjr=Δδr,1f_{j}^{r}=g_{j}^{r}=\Delta\delta_{r,1} and iTTiT\rightarrow T, the model as a paradigm of pp wave topological superconductivity has been well studied. Here, we take the imaginary hopping strength for the sake of convenience in the following discussion, and iTTiT\rightarrow T can be realized by local gauge transformation. However, the origin of the phase in the hopping cannot be considered as the magnetic flux since it still takes effect on the eigenstates even if the open boundary condition is taken due to the existence of the pair term. One of the possible origins is the phase gradient on the pair term seradjeh2011unpaired ; romito2012manipulating ; rontynen2014tuning ; dmytruk2019majorana ; takasan2022supercurrent .

The main purposes of the following discussion are (i) to present a Hermitian system with a superconducting ground state, which possesses the maximal BCS-pair order parameter, and (ii) to provide a scheme to achieve such a state dynamically by a non-Hermitian system with local pair source and drain. To this end, we will consider several types of functions {fjr,gjr}\left\{f_{j}^{r},g_{j}^{r}\right\}, which correspond to different types of pairing processes, including local and extensive sized pairing. In the case with spatially separated creation and annihilation of pairs, the non-Hermitian term is naturally involved, and then some particular dynamic behaviors may emerge that never appear in a Hermitian system. One of them is the EP dynamics, which provides a mechanism for a relaxation process in the framework of quantum mechanics. In fact, in the case with fjr=gjr=Δδr,1f_{j}^{r}=g_{j}^{r}=\Delta\delta_{r,1}, the model has been studied systematically Off-diagonal2022ma . It has been shown that the ground states near the critical point μ=0\mu=0 possess ODLRO in association with the maximum of BCS-pair order parameter.

In the framework of the Kitaev model, the pair number is not suitable for characterizing a superconducting state, since the fully filled pair state k>0ckck|0\prod_{k>0}c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\left|0\right\rangle =eiθl=1Ncl|0=e^{i\theta}\prod_{l=1}^{N}c_{l}^{{\dagger}}\left|0\right\rangle is an insulating state. To quantitatively characterize the superconductivity of a given state |ψ\left|\psi\right\rangle, we introduce the operator

𝒪=2Nk>0|ψ|ckck|ψ|.\mathcal{O}=\frac{2}{N}\sum_{k>0}|\left\langle\psi\right|c_{k}c_{-k}\left|\psi\right\rangle|. (4)

Obviously, for a given state |ψ\left|\psi\right\rangle, quantity |ψ|ckck|ψ||\left\langle\psi\right|c_{k}c_{-k}\left|\psi\right\rangle| =|ψ|ckck|ψ|=|\left\langle\psi\right|c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\left|\psi\right\rangle| measures the rate of transition for a pair at kk channel and the population of pairs. Then, 𝒪\mathcal{O} is defined by the average magnitude over all channels. In general, nonzero 𝒪\mathcal{O} means that state |ψ\left|\psi\right\rangle is a superconducting state.

III Perfect superconducting state

In this section, we consider a specific distribution function

D(r)=2iπΔrδr,odd,D(r)=-\frac{2i}{\pi}\frac{\Delta}{r}\delta_{r,\mathrm{odd}}, (5)

which is the key of the following toy model. Such a toy model allows us to obtain an exact solution, which is ultimately related to the main goal of this work. To this end, we introduce a set of pseudo spin operators

s\displaystyle s^{-} =\displaystyle= (s+)=k>0ckck,\displaystyle\left(s^{+}\right)^{{\dagger}}=\sum\limits_{k>0}c_{k}c_{-k}, (6)
sz\displaystyle s^{z} =\displaystyle= 12k>0(ckck+ckck1),\displaystyle\frac{1}{2}\sum\limits_{k>0}\left(c_{k}^{{\dagger}}c_{k}+c_{-k}^{{\dagger}}c_{-k}-1\right), (7)

which obey the SU(2) commutation relation, [s+,s]=2sz\left[s^{+},s^{-}\right]=2s^{z}. Particularly, in real space, the spin operator has the form

s=2iπjodd r1rcjcj+r,s^{-}=\frac{2i}{\pi}\sum\limits_{j}\sum\limits_{\mathrm{odd}\text{ }r}\frac{1}{r}c_{j}c_{j+r}, (8)

in the thermodynamic limit NN\longrightarrow\infty. Then, taking fjr=(gjr)=D(r)f_{j}^{r}=\left(g_{j}^{r}\right)^{\ast}=D^{\ast}(r), we have

HP=2Δiπjodd r1r(cjcj+rcj+rcj),H_{\mathrm{P}}=\frac{2\Delta i}{\pi}\sum\limits_{j}\sum\limits_{\mathrm{odd}\text{ }r}\frac{1}{r}\left(c_{j}^{{\dagger}}c_{j+r}^{{\dagger}}-c_{j+r}c_{j}\right), (9)

and the Hermitian extended Kitaev Hamiltonian in the form

Hext\displaystyle H_{\mathrm{ext}} =\displaystyle= 2Tk>0sink(ckckckck)+2μsz\displaystyle 2T\sum\limits_{k>0}\sin k(c_{k}^{{\dagger}}c_{k}-c_{-k}^{{\dagger}}c_{-k})+2\mu s^{z} (10)
+Δ(s++s).\displaystyle+\Delta\left(s^{+}+s^{-}\right).

Here, we neglect a constant μN/2\mu N/2 for the sake of simplicity. In the BCS-pair invariant subspace spanned by the pair states

|Ψ{k}={k}ckck|0,\left|\Psi_{\left\{k\right\}}\right\rangle=\prod_{\{k\}}c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\left|0\right\rangle, (11)

where {k}\left\{k\right\} denotes the 2N/22^{N/2} dimensional set of configuration of the BCS-pair filling, we always have (ckckckck)|Ψ{k}(c_{k}^{{\dagger}}c_{k}-c_{-k}^{{\dagger}}c_{-k})\left|\Psi_{\left\{k\right\}}\right\rangle =0=0\ and (ckckckck)|0(c_{k}^{{\dagger}}c_{k}-c_{-k}^{{\dagger}}c_{-k})\left|0\right\rangle =0=0. This indicates that all 2N/22^{N/2} states {|Ψ{k}}\left\{\left|\Psi_{\left\{k\right\}}\right\rangle\right\} are zero-energy eigenstates of HTH_{\mathrm{T}}, i.e., the first term of HextH_{\mathrm{ext}}. Then, we have the equivalent Hamiltonian of HextH_{\mathrm{ext}} in the subspace

ext=𝐁𝐬,\mathcal{H}_{\mathrm{ext}}=\mathbf{B}\cdot\mathbf{s}, (12)

where the magnetic field 𝐁=(2Δ,0,2μ)\mathbf{B=}(2\Delta,0,2\mu). Obviously, both 𝐬2\mathbf{s}^{2} and szs^{z} are commutative to HextH_{\mathrm{ext}}. Based on this fact, one can further construct multi-invariant subspaces by the common eigenstates of 𝐬2\mathbf{s}^{2} and szs^{z}. We are interested in a set {|ψn}\left\{\left|\psi_{n}\right\rangle\right\} with n[0,N/2]n\in\left[0,N/2\right]

|ψn=1Ωn(s+)n|0,\left|\psi_{n}\right\rangle=\frac{1}{\Omega_{n}}(s^{+})^{n}\left|0\right\rangle, (13)

which obeys 𝐬2|ψn=\mathbf{s}^{2}\left|\psi_{n}\right\rangle= N(N/4+1)/4|ψnN\left(N/4+1\right)/4\left|\psi_{n}\right\rangle and sz|ψn=s^{z}\left|\psi_{n}\right\rangle= (nN/4)|ψn\left(n-N/4\right)\left|\psi_{n}\right\rangle, with the normalization factor Ωn=\Omega_{n}= (n!)CN/2n\left(n!\right)\sqrt{C_{N/2}^{n}}. The eigenstates of HextH_{\mathrm{ext}} are actually the eigenstates of the spin operator

𝐁|𝐁|𝐬=1Δ2+μ2(Δsx+μsz),\frac{\mathbf{B}}{\left|\mathbf{B}\right|}\cdot\mathbf{s=}\frac{1}{\sqrt{\Delta^{2}+\mu^{2}}}(\Delta s^{x}+\mu s^{z}), (14)

and be expressed in the form

|Φl(μ)=ndnl(μ)|ψn,\left|\Phi_{l}(\mu)\right\rangle=\sum_{n}d_{n}^{l}(\mu)\left|\psi_{n}\right\rangle, (15)

satisfying the equation

Hext|Φl(μ)=2Δ2+μ2(lN4)|Φl(μ).H_{\mathrm{ext}}\left|\Phi_{l}(\mu)\right\rangle=2\sqrt{\Delta^{2}+\mu^{2}}\left(l-\frac{N}{4}\right)\left|\Phi_{l}(\mu)\right\rangle. (16)

The coefficient dnl(μ)d_{n}^{l}(\mu) can be obtained exactly, but here, we only list two of them explicitly

dn0(μ)=CN/2n(1)ncosnδ2sin(N/2n)δ2,\displaystyle d_{n}^{0}(\mu)=\sqrt{C_{N/2}^{n}}\left(-1\right)^{n}\cos^{n}\frac{\delta}{2}\sin^{\left(N/2-n\right)}\frac{\delta}{2}, (17)
dnN/2(μ)=CN/2nsinnδ2cos(N/2n)δ2,\displaystyle d_{n}^{N/2}(\mu)=\sqrt{C_{N/2}^{n}}\sin^{n}\frac{\delta}{2}\cos^{\left(N/2-n\right)}\frac{\delta}{2}, (18)

which lead to

|Φ0(μ)=k>0(sinδ2cosδ2ckck)|0,\displaystyle\left|\Phi_{0}(\mu)\right\rangle=\prod\limits_{k>0}\left(\sin\frac{\delta}{2}-\cos\frac{\delta}{2}c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\right)\left|0\right\rangle, (19)
|ΦN/2(μ)=k>0(cosδ2+sinδ2ckck)|0,\displaystyle\left|\Phi_{N/2}(\mu)\right\rangle=\prod\limits_{k>0}\left(\cos\frac{\delta}{2}+\sin\frac{\delta}{2}c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\right)\left|0\right\rangle, (20)

with tanδ=Δ/μ\tan\delta=-\Delta/\mu. We note that the two above eigenstates reduce to k>0(1\prod_{k>0}(1\mp ckck)c_{-k}^{{\dagger}}c_{k}^{{\dagger}}) /2|0k|0k/\sqrt{2}\left|0\right\rangle_{k}\left|0\right\rangle_{-k}, which supports that the corresponding order parameter reaches the maximum 0.50.5 when taking the chemical potential μ=0\mu=0. We refer to such states as perfect superconducting states.

It is clear that these two states are not unique perfect superconducting states. In fact, state k>0(1+\prod_{k>0}(1+ eiγkckck)e^{i\gamma_{k}}c_{-k}^{{\dagger}}c_{k}^{{\dagger}}) /2|0/\sqrt{2}\left|0\right\rangle has the same feature for any distribution of {γk}\left\{\gamma_{k}\right\}. This inspires us to find another way to prepare a superconducting state. Now we consider the dynamic generation of such states in the case with zero μ\mu. For the initial state |ψ(0)=|ψ0\left|\psi\left(0\right)\right\rangle=\left|\psi_{0}\right\rangle, which is actually an empty state, the time evolution under a quench Hamiltonian

Hquen=Δ(s+s+),H_{\mathrm{quen}}=\Delta(s^{-}+s^{+}), (21)

can be expressed as

|ψ(t)=exp[iΔ(s+s+)t]|ψ(0),\left|\psi\left(t\right)\right\rangle=\exp\left[-i\Delta(s^{-}+s^{+})t\right]\left|\psi\left(0\right)\right\rangle, (22)

which is essentially a rotation around the xx-axis. Obviously |ψ(t)\left|\psi\left(t\right)\right\rangle can be easily obtained by local rotation for each kk,

|ψ(t)=k>0(cosθisinθckck)|0,\left|\psi\left(t\right)\right\rangle=\prod\limits_{k>0}\left(\cos\theta-i\sin\theta c_{-k}^{{\dagger}}c_{k}^{{\dagger}}\right)\left|0\right\rangle, (23)

where θ\theta is a function of time θ(t)=Δt\theta(t)=\Delta t. Direct derivation shows that the order parameter is a periodic function of time

𝒪(t)=12|sin(2Δt)|,\mathcal{O}(t)=\frac{1}{2}\left|\sin(2\Delta t)\right|, (24)

which reaches 0.50.5 at instants t=(m+1/2)π/(2Δ)t=(m+1/2)\pi/(2\Delta) with integer mm. This indicates that a perfect superconducting state can be established via a dynamic process. The physics seems to be clear that the oscillating 𝒪(t)\mathcal{O}(t) is a resultant effect of both pair creation and annihilation terms.

IV High-order EP and dynamics

Now, we consider a question of what happens if only the pair annihilation (creation) terms are taken. It is a first step to investigate the effect of spatially separated source and drain. Naturally, a non-Hermitian Hamiltonian by taking fjr=0f_{j}^{r}=0 but gjr=D(r)g_{j}^{r}=D(r) is involved, i.e.,

HP=2iΔπj=1Nodd r1rcj+rcj.H_{\mathrm{P}}=-\frac{2i\Delta}{\pi}\sum\limits_{j=1}^{N}\sum\limits_{\mathrm{odd}\text{ }r}\frac{1}{r}c_{j+r}c_{j}. (25)

The equivalent Hamiltonian in the invariant subspace spanned by the set of states {|ψn}\left\{\left|\psi_{n}\right\rangle\right\} becomes

ext=Δs.\mathcal{H}_{\mathrm{ext}}=\Delta s^{-}. (26)

The dynamics of ext\mathcal{H}_{\mathrm{ext}} are slightly little special and can be captured from the matrix representation of Hamiltonian ext\mathcal{H}_{\mathrm{ext}}. It is an (N/2+1)×(N/2+1)\left(N/2+1\right)\times\left(N/2+1\right) matrix MM, with nonzero matrix elements

(M)N/2n,N/2+1n=Δn(N/2n+1),\left(M\right)_{N/2-n,N/2+1-n}=\Delta\sqrt{n\left(N/2-n+1\right)}\text{,} (27)

with n=[0,N/21]n=\left[0,N/2-1\right]. Note that MM is a nilpotent matrix, i.e. MN/2+1=0M^{N/2+1}=0, or an (N/2+1)\left(N/2+1\right)-order Jordan block. The dynamics for any state in this subspace {|ψn}\left\{\left|\psi_{n}\right\rangle\right\} is governed by the time evolution operator

U(t)=eiMt=l=0N/21l!(iMt)l.U(t)=e^{-iMt}=\sum_{l=0}^{N/2}\frac{1}{l!}\left(-iMt\right)^{l}. (28)

Then for the initial state |ψ(0)=|ψN/2\left|\psi\left(0\right)\right\rangle=\left|\psi_{N/2}\right\rangle, we have the normalized evolved state

|ψ(t)=k>0itΔ+ckck1+Δ2t2|0,\left|\psi\left(t\right)\right\rangle=\prod_{k>0}\frac{-it\Delta+c_{-k}^{{\dagger}}c_{k}^{{\dagger}}}{\sqrt{1+\Delta^{2}t^{2}}}\left|0\right\rangle, (29)

which turns to the coalescing state, i.e., |ψ()|ψ0\left|\psi\left(\infty\right)\right\rangle\longrightarrow\left|\psi_{0}\right\rangle. Obviously, |ψ(t)\left|\psi\left(t\right)\right\rangle has maximal 𝒪\mathcal{O} at instant 1/Δ1/\Delta. The above analysis is still true when we take fjr=D(r)f_{j}^{r}=D^{\ast}(r) but gjr=0g_{j}^{r}=0 and |ψ(0)=|ψ0\left|\psi\left(0\right)\right\rangle=\left|\psi_{0}\right\rangle, which corresponds to a time reversal process. As expected, the physical picture is clear that the pair term takes the role of not only pair generation but also reduction. Intuitively, a local pair term should have a similar effect. This is the aim of the next section.

V Single source or drain

The results obtained in the last section are exact and explicit due to the translational symmetry of the model. In this section, we will show that a similar result can be obtained approximately when only a single pair term is considered, i.e., fjr=0f_{j}^{r}=0 but gjr=δj,j0D(r)g_{j}^{r}=\delta_{j,j_{0}}D(r), or vice versa. First, states {|ψn}\left\{\left|\psi_{n}\right\rangle\right\} have translational symmetry with zero momentum. This originates from the translational symmetry of the system, i. e., [HT,𝒯1]=0\left[H_{\mathrm{T}},\mathcal{T}_{1}\right]=0, where operator 𝒯1\mathcal{T}_{1} is defined by 𝒯1cj𝒯11=cj+1\mathcal{T}_{1}c_{j}\mathcal{T}_{1}^{-1}=c_{j+1}. Obviously, we have 𝒯1|ψ0=|ψ0\mathcal{T}_{1}\left|\psi_{0}\right\rangle=\left|\psi_{0}\right\rangle, which results in

𝒯1|ψn=α|ψn,\mathcal{T}_{1}\left|\psi_{n}\right\rangle=\alpha\left|\psi_{n}\right\rangle, (30)

with |α|=1\left|\alpha\right|=1 because 𝒯11s±𝒯1=s±\mathcal{T}_{1}^{-1}s^{\pm}\mathcal{T}_{1}=s^{\pm}. Then, we have if NN\longrightarrow\infty

ψn|odd r1rcj+rcj|ψm=πi2Nm(N/2m+1)δn,m1,\left\langle\psi_{n}\right|\sum\limits_{\mathrm{odd}\text{ }r}\frac{1}{r}c_{j+r}c_{j}\left|\psi_{m}\right\rangle=\frac{\pi i}{2N}\sqrt{m\left(N/2-m+1\right)}\delta_{n,m-1}, (31)

based on the relation

ψn|cj+rcj|ψm=ψn|cj+r+1cj+1|ψm.\left\langle\psi_{n}\right|c_{j+r}c_{j}\left|\psi_{m}\right\rangle=\left\langle\psi_{n}\right|c_{j+r+1}c_{j+1}\left|\psi_{m}\right\rangle. (32)

Obviously, the perturbation matrix is still in (N+1N+1)-order Jordan block form. The time evolution under such a system should obey the EP dynamics.

Furthermore, a Jordan block matrix does not restrict the values of the nonzero matrix elements. Then the relation

ψn|cj+1cj|ψmδm,n+1,\left\langle\psi_{n}\right|c_{j+1}c_{j}\left|\psi_{m}\right\rangle\propto\delta_{m,n+1}, (33)

may also result in EP dynamics, based on the following analysis. Actually, considering a more generalized form of HP=i,jλijcicjH_{\mathrm{P}}=\sum_{i,j}\lambda_{ij}c_{i}c_{j} with arbitrary factor {λij}\left\{\lambda_{ij}\right\}, states |ψ0\left|\psi_{0}\right\rangle and |ψN\left|\psi_{N}\right\rangle are two degenerate states of the Hermitian Hamiltonian HTH_{\mathrm{T}}, and we always have

|ψ0=0,|ψN/2=0,\mathcal{H}\left|\psi_{0}\right\rangle=0,\mathcal{H}^{{\dagger}}\left|\psi_{N/2}\right\rangle=0, (34)

due to the facts

cj+rcj|ψ0=0,(cj+rcj)|ψN/2=0.c_{j+r}c_{j}\left|\psi_{0}\right\rangle=0,\left(c_{j+r}c_{j}\right)^{{\dagger}}\left|\psi_{N/2}\right\rangle=0. (35)

This means that two states |ψ0\left|\psi_{0}\right\rangle and that |ψN/2\left|\psi_{N/2}\right\rangle are mutually biorthogonal conjugates and ψ0|ψN/2\langle\psi_{0}\left|\psi_{N/2}\right\rangle is their biorthogonal norm. Importantly, the vanishing norm ψ0|ψN/2=0\langle\psi_{0}\left|\psi_{N/2}\right\rangle=0 indicates that state |ψ0\left|\psi_{0}\right\rangle(|ψN/2\left|\psi_{N/2}\right\rangle) is the coalescing state of HH(HH^{{\dagger}}) or Hamiltonians HH and HH^{{\dagger}} obtain an EP. From the perspective of dynamics, we have

eiHt|ψN/2|ψ0,eiHt|ψ0|ψN/2,e^{-iHt}\left|\psi_{N/2}\right\rangle\longrightarrow\left|\psi_{0}\right\rangle,e^{-iH^{{\dagger}}t}\left|\psi_{0}\right\rangle\longrightarrow\left|\psi_{N/2}\right\rangle, (36)

for a sufficiently long time tt. Although both states |ψ0\left|\psi_{0}\right\rangle and |ψN/2\left|\psi_{N/2}\right\rangle are trivial states, eiHt|ψN/2e^{-iHt}\left|\psi_{N/2}\right\rangle\ and eiHt|ψ0e^{-iH^{{\dagger}}t}\left|\psi_{0}\right\rangle may have pair currents at finite tt.

Refer to caption
Figure 2: Plots of the time evolution of 𝒪\mathcal{O} in Eq. (43) for several representative φ\varphi under the Hamiltonian with HTH_{\mathrm{T}} in Eq. (2) and HquenH_{\mathrm{quen}} in Eq. (42) on the lattice, which is schematically illustrated in Fig. 1 (a) and (b), respectively. The initial state is the vacuum state and the parameters are N=9N=9, T=1T=1, μ=0\mu=0 and Δ=0.005\Delta=0.005. (a1), (a2) and (a3) are the situations for a ring lattice with N0=N_{0}= 33, 44 and 55, respectively. (b1), (b2) and (b3) are the situations for an open chain with N0=N_{0}= 33, 55 and 88, respectively. We find that for all cases, 𝒪(t)\mathcal{O}(t) tends to stabilize at 𝒪()\mathcal{O}(\infty) after a sufficiently long time. The results show that the stable final state has maximal 𝒪\mathcal{O} for different φ\varphi. 𝒪()\mathcal{O}(\infty) is not sensitive to N0N_{0} for the ring system, and φ\varphi affects the rate of convergence of the evolved state. For the chain system, 𝒪()\mathcal{O}(\infty) reaches the maximum when the pair source and drain are located at the ends of the chain. 

VI Resonant distant source and drain

The above result at least indicates that the local pair terms can be regarded as particle sources or drains, which can fully fill the empty state (|ψ0|ψN/2\left|\psi_{0}\right\rangle\longrightarrow\left|\psi_{N/2}\right\rangle) or empty the fully filled state (|ψN/2|ψ0\left|\psi_{N/2}\right\rangle\longrightarrow\left|\psi_{0}\right\rangle). This inspires us to investigate the dynamics with balanced local pair terms that are spatially separated by a distance. To this end, we consider the pair term in the resonant form

fjr=eiφδj,1D(r),gjr=δj,N0D(r),f_{j}^{r}=e^{i\varphi}\delta_{j,1}D^{\ast}(r),g_{j}^{r}=\delta_{j,N_{0}}D(r), (37)

or explicitly

HP=2iΔπodd,r1r(eiφc1c1+rcN0+rcN0),H_{\mathrm{P}}=\frac{2i\Delta}{\pi}\sum\limits_{\mathrm{odd},r}\frac{1}{r}\left(e^{i\varphi}c_{1}^{{\dagger}}c_{1+r}^{{\dagger}}-c_{N_{0}+r}c_{N_{0}}\right), (38)

which acts as separated local pair sources and drains. Here, as the resonant condition, the amplitudes of the pair annihilation and creation terms are the same, while there is a phase difference φ\varphi between them, which is crucial for the dynamics, as shown in the following. In the small Δ\Delta limit, based on the perturbation method, the matrix representation of Hamiltonian ext\mathcal{H}_{\mathrm{ext}} in the subspace spanned by the set of states {einφ/2|ψn}\left\{e^{in\varphi/2}\left|\psi_{n}\right\rangle\right\} is an (N/2+1)×(N/2+1)\left(N/2+1\right)\times\left(N/2+1\right) matrix ext\mathcal{H}_{\mathrm{ext}} with nonzero matrix elements

(ext)N/2+1n,N/2n\displaystyle\left(\mathcal{H}_{\mathrm{ext}}\right)_{N/2+1-n,N/2-n} =\displaystyle= eiφ/2ΔNn(N/2+1n)\displaystyle e^{i\varphi/2}\frac{\Delta}{N}\sqrt{n\left(N/2+1-n\right)} (39)
=\displaystyle= (ext)N/2n,N/2+1n,\displaystyle\left(\mathcal{H}_{\mathrm{ext}}\right)_{N/2-n,N/2+1-n},

with n=[0,N/21]n=\left[0,N/2-1\right]. We note that matrix ext\mathcal{H}_{\mathrm{ext}} is the same as that in (21) but with a complex strength constant. The corresponding eigenenergy is complex

En=Δeiφ/2N(2nN2),E_{n}=\frac{\Delta e^{i\varphi/2}}{N}(2n-\frac{N}{2}), (40)

with n=[0,N/2]n=\left[0,N/2\right], and its imaginary part is Im(En)=\mathrm{Im}\left(E_{n}\right)= Δ(2n/N1/2)sinφ2\Delta\left(2n/N-1/2\right)\sin\frac{\varphi}{2}. Unlike a Hermitian system, the imaginary part of the eigenvalue can amplify or reduce the corresponding amplitude of the wave function in the dynamic process. For the given initial state |ψ(0)=|ψ0\left|\psi\left(0\right)\right\rangle=\left|\psi_{0}\right\rangle when the evolution time is sufficiently long, the final state is the eigenstate of ext\mathcal{H}_{\mathrm{ext}} with the maximum imaginary part. The corresponding approximate eigenstate is

|ψ()=k>0σeiφ/2+ckck2|0,\left|\psi\left(\infty\right)\right\rangle=\prod\limits_{k>0}\frac{\sigma e^{-i\varphi/2}+c_{-k}^{{\dagger}}c_{k}^{{\dagger}}}{\sqrt{2}}\left|0\right\rangle, (41)

where σ=sgn(sinφ2)\sigma=\mathrm{sgn}(\sin\frac{\varphi}{2}). Obviously, |ψ()\left|\psi\left(\infty\right)\right\rangle is a perfect superconducting state. When the off-resonant case is considered, the expression is the same as |ψ()\left|\psi\left(\infty\right)\right\rangle, but an imaginary part should be added in φ\varphi, which will reduce the order parameter from 0.50.5.

Now we consider a more practical case with

quen=Δ(eiφc2c1+cN0cN0+1),\mathcal{H}_{\mathrm{quen}}=\Delta\left(e^{i\varphi}c_{2}^{{\dagger}}c_{1}^{{\dagger}}+c_{N_{0}}c_{N_{0}+1}\right), (42)

where the pairing terms reduce to the simplest case. In addition, we also consider the case with an open boundary condition, which is closer to the real sample in the experiment. The physical intuition for this setup is simple. Term c1c2c_{1}^{{\dagger}}c_{2}^{{\dagger}} acts as a source of pair at one end of the chain, while cN1cNc_{N-1}c_{N} takes the role of drain at the other end. According to the analysis of the pair term i,jλijcicj\sum_{i,j}\lambda_{ij}c_{i}c_{j} in the last section, it is expected that the nearest neighboring pair terms share a similar feature, i.e., a stable state with the order parameter close to that of state (41) emerges when the source and drain are balanced. Numerical simulation is performed to verify our predictions. We compute the time evolution |ψ(t)=ei(HT+quen)t|ψ(0)\left|\psi\left(t\right)\right\rangle=e^{-i\left(H_{\mathrm{T}}+\mathcal{H}_{\mathrm{quen}}\right)t}\left|\psi\left(0\right)\right\rangle by exact diagonalization. The geometries of finite systems are schematically illustrated in Fig. 1. We consider an N=9N=9 site system with periodic and open boundary conditions. In this case, the order parameter has the following explicit form:

𝒪(t)=14n=14|ψ(t)|c2πn9c2πn9|ψ(t)|ψ(t)|ψ(t).\mathcal{O}(t)=\frac{1}{4}\sum_{n=1}^{4}\frac{|\left\langle\psi\left(t\right)\right|c_{\frac{2\pi n}{9}}c_{-\frac{2\pi n}{9}}\left|\psi\left(t\right)\right\rangle|}{\left\langle\psi\left(t\right)\right|\psi\left(t\right)\rangle}. (43)

We plot 𝒪(t)\mathcal{O}(t) as a function of tt and φ\varphi in Fig. 2 for the finite size cases schematically illustrated in Fig. 1, obtained by numerical simulations. The numerical results agree with our prediction, that for all cases, 𝒪(t)\mathcal{O}(t) tends to stabilize at 𝒪()\mathcal{O}(\infty) after a sufficiently long time. In addition, we find that the relaxation process and the final 𝒪\mathcal{O} depend on the geometry and φ\varphi. (i) 𝒪\mathcal{O}()(\infty) is not sensitive to N0N_{0} for the ring system, and 𝒪\mathcal{O}()(\infty) can reach the maximum 0.50.5. (ii) 𝒪\mathcal{O}()(\infty) depends on N0N_{0} for the chain system, and 𝒪\mathcal{O}()(\infty) can reach the maximum 0.450.45 when the pair source and drain are located at the ends of the chain. This implies that the balanced edge pair source and drain benefit to forming a superconducting state. For both boundary conditions, the converging time depends on the value of φ\varphi, but in two different ways.

VII Summary

In summary, we have studied several types of toy models with deliberately engineered pair terms to explore the possibility of realizing nonequilibrium superconducting state in a nonhomogeneous Kitaev model, which is essentially a non-Hermitian extension of the Kitaev chain. In the framework of quantum mechanics, based on the analysis of the exact solution and perturbation method, we find that the EP dynamics provides a clear picture for the action of a single pair source or drain. When the two coexist simultaneously, the dynamics depend on the distance and the relative phase between the pair creation and annihilation terms. Analytical analysis and numerical simulation results show that the superconducting state can be dynamically established at the resonant local pair source and drain. Two spatially separated pair terms can drive an initial empty state to a stationary state with a near maximal pair order parameter. It provides an alternative way of understanding the mechanism of the nonequilibrium superconducting state.

Acknowledgement

This work was supported by the National Natural Science Foundation of China (under Grant No. 11874225).

References

  • (1) T. Kinoshita, T. Wenger, and D. S. Weiss, A quantum Newton’ cradle, Nature (London) 440, 900 (2006).
  • (2) S. Trotzky, Y.-A. Chen, A. Flesch, I. P. McCulloch, U. Schollwöck, J. Eisert, and I. Bloch, Probing the relaxation towards equilibrium in an isolated strongly correlated 1D Bose gas, Nat. Phys. 8, 325 (2012).
  • (3) M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, B. Rauer, M. Schreitl, I. Mazets, D. A. Smith, E. Demler, and J. Schmiedmayer, Relaxation and Prethermalization in an Isolated Quantum System, Science 337, 1318 (2012).
  • (4) M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Lüschen, M. H. Fischer, R. Vosk, E. Altman, U. Schneider, and I. Bloch, Observation of many-body localization of interacting fermions in a quasirandom optical lattice, Science 349, 842 (2015).
  • (5) J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess, P. Hauke, M. Heyl, D. A. Huse, and C. R. Monroe, Many Body Localization in a Quantum Simulator with Programmable Random Disorder, Nat. Phys. 12, 907 (2016).
  • (6) A. M. Kaufman, M. E. Tai, A. Lukin, M. Rispoli, R. Schittko, P. M. Preiss, and M. Greiner, Quantum Thermalization through Entanglement in an Isolated Many-Body System, Science 353, 794 (2016).
  • (7) X. Z. Zhang and Z. Song, Steady off-diagonal long-range order state in a half-filled dimerized Hubbard chain induced by a resonant pulsed field, Phys. Rev. B 106, 094301 (2022).
  • (8) S. Choi, J. Choi, R. Landig, G. Kucsko, H. Zhou, J. Isoya, F. Jelezko, S. Onoda, H. Sumiya, V. Khemani, C. v. Keyserlingk, N. Y. Yao, E. Demler, and M. D. Lukin, Observation of discrete time-crystalline order in a disordered dipolar many-body system, Nature 543, 221 (2017).
  • (9) D. V. Else, B. Bauer, and C. Nayak, Floquet time crystals, Phys. Rev. Lett. 117, 090402 (2016).
  • (10) V. Khemani, A. Lazarides, R. Moessner, and S. L. Sondhi, Phase structure of driven quantum systems, Phys. Rev. Lett. 116, 250401 (2016).
  • (11) N. H. Lindner, G. Refael, and V. Galitski, Floquet Topological Insulator in Semiconductor Quantum Wells, Nat. Phys. 7, 490 (2011).
  • (12) T. Kaneko, T. Shirakawa, S. Sorella, and S. Yunoki, Photoinduced η\eta Pairing in the Hubbard Model, Phys. Rev. Lett.  122, 077002 (2019).
  • (13) J. Tindall, B. Buča, J. R. Coulthard, and D. Jaksch, Heating-Induced Long-Range  η\eta Pairing in the Hubbard Model, Phys. Rev. Lett. 123, 030603 (2019).
  • (14) X. M. Yang and Z. Song, Resonant generation of a p-wave Cooper pair in a non-Hermitian Kitaev chain at the exceptional point, Phys. Rev. A 102, 022219 (2020).
  • (15) M. Rigol, V. Dunjko, and M. Olshanii, Thermalization and Its Mechanism for Generic Isolated Quantum Systems, Nature (London) 452, 854 (2008).
  • (16) J. Eisert, M. Friesdorf, and C. Gogolin, Quantum Many-Body Systems out of Equilibrium, Nat. Phys. 11, 124 (2015).
  • (17) M. Heyl, Dynamical quantum phase transitions: A brief survey, Europhys. Lett. 125, 26001 (2019).
  • (18) S. Kitamura and H. Aoki, η\eta-pairing superfluid in periodically-driven fermionic Hubbard model with strong attraction, Phys. Rev. B 94, 174503 (2016).
  • (19) T. Kaneko, T. Shirakawa, S. Sorella, and S. Yunoki, Photoinduced η\eta Pairing in the Hubbard Model, Phys. Rev. Lett. 122, 077002 (2019).
  • (20) J. Tindall, B. Buča, J. R. Coulthard, and D. Jaksch, Heating-Induced Long-Range η\eta Pairing in the Hubbard Model, Phys. Rev. Lett. 123, 030603 (2019).
  • (21) R. Fujiuchi, T. Kaneko, Y. Ohta, and S. Yunoki, Photoinduced electron-electron pairing in the extended Falicov-Kimball model, Phys. Rev. B 100, 045121 (2019).
  • (22) F. Peronaci, O. Parcollet, and M. Schiró, Enhancement of local pairing correlations in periodically driven Mott insulators, Phys. Rev. B 101, 161101 (2020).
  • (23) R. Fujiuchi, T. Kaneko, K. Sugimoto, S. Yunoki, and Y. Ohta, Superconductivity and charge density wave under a time-dependent periodic field in the one-dimensional attractive Hubbard model, Phys. Rev. B 101, 235122 (2020).
  • (24) S. Ejima, T. Kaneko, F. Lange, S. Yunoki, and H. Fehske, Photoinduced η\eta-pairing at finite temperatures, Phys. Rev. Research 2, 032008(R) (2020).
  • (25) J. Li, D. Golez, P. Werner, and M. Eckstein, η\eta-Paired Superconducting Hidden Phase in Photodoped Mott Insulators, Phys. Rev. B 102, 165136 (2020).
  • (26) T. Kaneko, S. Yunoki, and A. J. Millis, Charge stiffness and long-range correlation in the optically induced η\eta-pairing state of the one-dimensional Hubbard model, Phys. Rev. Research 2, 032027(R) (2020).
  • (27) X. Z. Zhang and Z. Song, η\eta-pairing ground states in the non-Hermitian Hubbard model, Phys. Rev. B 103, 235153 (2021).
  • (28) X. Z. Zhang and Z. Song, Dynamical preparation of a steady off-diagonal long-range order state in the Hubbard model with a local non-Hermitian impurity, Phys. Rev. B 102, 174303 (2020).
  • (29) X. M. Yang and Z. Song, Dynamic transition from insulating state to η\eta-pairing state in a composite non-Hermitian system, Phys. Rev. B 105, 195132 (2022).
  • (30) C. N. Yang, η\eta Pairing and Off-Diagonal Long-Range Order in a Hubbard Model, Phys. Rev. Lett. 63, 2144 (1989).
  • (31) A. Y. Kitaev, Unpaired Majorana fermions in quantum wires, Phys. Usp. 44, 131 (2001).
  • (32) C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. D. Sarma, Non-Abelian anyons and topological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).
  • (33) A. Stern, Non-Abelian states of matter, Nature (London) 464, 187 (2010).
  • (34) J. Alicea, New directions in the pursuit of Majorana fermions in solid state systems, Rep. Prog. Phys. 75, 076501 (2012).
  • (35) Y. B. Shi , K. L. Zhang , and Z. Song, Dynamic generation of nonequilibrium superconducting states in a Kitaev chain, Phys. Rev. B 106, 184505 (2022).
  • (36) C. M. Bender and S. Boettcher, Real spectra in non-Hermitian Hamiltonians having PT symmetry, Phys. Rev. Lett. 80, 5243 (1998).
  • (37) C. M. Bender, D. C. Brody, and H. F. Jones, Complex extension of quantum mechanics, Phys. Rev. Lett. 89, 270401 (2002).
  • (38) X. J. Liu, C. L. M. Wong, and K. T. Law, Non-abelian majorana doublets in time-reversal-invariant topological superconductors, Phys. Rev. X 4, 021018 (2014).
  • (39) X. H. Wang, T. T. Liu, Y. Xiong, and P. Q. Tong, Spontaneous PT-symmetry breaking in non-Hermitian Kitaev and extended Kitaev models, Phys. Rev. A 92, 012116 (2015).
  • (40) C. Yuce, Majorana edge modes with gain and loss, Phys. Rev. A 93, 062130 (2016).
  • (41) Q. B. Zeng, B. G. Zhu, S. Chen, L. You, and R. Lü, Non-Hermitian Kitaev chain with complex on-site potentials, Phys. Rev. A 94, 022119 (2016).
  • (42) M. Klett, H. Cartarius, D. Dast, J. Main, and G. Wunner, Relation between PT-symmetry breaking and topologically nontrivial phases in the Su-Schrieffer-Heeger and Kitaev models, Phys. Rev. A 95, 053626 (2017).
  • (43) H. Menke and M. M. Hirschmann, Topological quantum wires with balanced gain and loss, Phys. Rev. B 95, 174506 (2017).
  • (44) C. Li, X. Z. Zhang, G. Zhang, and Z. Song, Topological phases in a Kitaev chain with imbalanced pairing, Phys. Rev. B 97, 115436 (2018).
  • (45) C. Li, G. Zhang, X. Z. Zhang, and Z. Song, Conventional quantum phase transition driven by a complex parameter in a non-Hermitian PT-symmetric Ising model, Phys. Rev. A 90, 012103 (2014).
  • (46) C. Li and Z. Song, Finite-temperature quantum criticality in a complex-parameter plane, Phys. Rev. A 92, 062103 (2015).
  • (47) C. Li, G. Zhang, and Z. Song, Chern number in Ising models with spatially modulated real and complex fields, Phys. Rev. A 94, 052113 (2016).
  • (48) X. Z. Zhang and Z. Song, Non-Hermitian anisotropic XY model with intrinsic rotation-time-reversal symmetry, Phys. Rev. A 87, 012114 (2013).
  • (49) M. V. Berry, Physics of nonhermitian degeneracies, Czech. J. Phys. 54, 1039 (2004).
  • (50) W. D. Heiss, The physics of exceptional points, J. Phys. A: Math. Theor. 45, 444016 (2012).
  • (51) M.-A. Miri and A. Alú, Exceptional points in optics and photonics, Science 363, eaar7709 (2019).
  • (52) X. Zhang and J. Gong, Non-Hermitian Floquet topological phases: Exceptional points, coalescent edge modes, and the skin effect, Phys. Rev. B 101, 045415 (2020).
  • (53) B. Seradjeh and E. Grosfeld, Unpaired Majorana fermions in a layered topological superconductor, Phys. Rev. B 83, 174521 (2011).
  • (54) A. Romito, J. Alicea, G. Refael, and F. von Oppen, Manipulating Majorana fermions using supercurrents, Phys. Rev. B 85, 020502(R) (2012).
  • (55) J. Röntynen and T. Ojanen, Tuning topological superconductivity in helical Shiba chains by supercurrent, Phys. Rev. B 90, 180503(R) (2014).
  • (56) O. Dmytruk, M. Thakurathi, D. Loss, and J. Klinovaja, Majorana bound states in double nanowires with reduced Zeeman thresholds due to supercurrents, Phys. Rev. B 99, 245416 (2019).
  • (57) K. Takasan, S. Sumita and Y. Yanase, Supercurrent-induced topological phase transitions, Phys. Rev. B 106, 014508 (2022).
  • (58) E. S. Ma and Z. Song, Off-diagonal long-range order in the ground state of the Kitaev chain, arXiv: 2212.02694.