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

Real-space decomposition of pp-wave Kitaev chain

D. K. He, E. S. Ma    Z. Song [email protected] School of Physics, Nankai University, Tianjin 300071, China School of Physics, Nankai University, Tianjin 300071, China
Abstract

We propose an extended Bogoliubov transformation in real space for spinless fermions, based on which a class of Kitaev chains of length 2N2N with zero chemical potential can be mapped to two independent Kitaev chains of length NN. It provides an alternative way to investigate a complicated system from the result of relatively simple systems. We demonstrate the implications of this decomposition by a Su-Schrieffer-Heeger (SSH) Kitaev model, which supports rich quantum phases. The features of the system, including the groundstate topology and nonequilibrium dynamics, can be revealed directly from that of sub-Kitaev chains. Based on this connection, two types of Bardeen-Cooper-Schrieffer (BCS)-pair order parameters are introduced to characterize the phase diagram, showing the ingredient of two different BCS pairing modes. Analytical analysis and numerical simulations show that the real-space decomposition for the ground state still holds true approximately in presence of finite chemical potential in the gapful regions.

I Introduction

The exact solution of a model Hamiltonian, especially for many-body system, plays an important role in physics and sometimes may open the door to the exploration of new frontiers in physics. One of common and efficient tools is the Fourier transformation, which decomposes a Hamiltonian into many commutative sub-Hamiltonians due to the translational symmetry. In contrast to the kk-space decomposition, there exist real-space decompositions, such as the block-diagonalization based on the reflection symmetry and so on. In this work, we propose another decomposition for the many-body Hamiltonian, which bases on an extended Bogoliubov transformations in real space for spinless fermions. By taking this transformation, we find that the Hamiltonians of a class of Kitaev chains Kitaev of length 2N2N with zero chemical potential can be mapped to two commutative ones of Kitaev chains of length NN. All the eigen states can be constructed by the ones of two sub-systems. It provides a clear physical picture and a simple way to solve the dynamic problem in a complicated system from the result for relative simple systems. We demonstrate the implications of this decomposition by a SSH Kitaev model, which consists of dimerized hopping and pairing terms and supports the rich quantum phases. The features of the system, including the groundstate topology and nonequilibrium dynamics from a trivial initial state, can be revealed directly from that of well-known simple Kitaev chains. Based on this connection, two types of BCS-pair order parameters are introduced to characterize the phase diagram, showing the ingredient of two different BCS pairing modes.

For instance, the topological index of the original model can be shown to be simply the sum of that of two sub-models. The ground state of an SSH Kitaev model has rich quantum phases with winding numbers 𝒩=0\mathcal{N=}0, 11, and 22, respectively, while a simple sub-system has quantum phases with winding numbers 𝒩=0\mathcal{N=}0 and 11. The physical picture becomes clear in the framework of decomposition. Based on the exact solution, the BCS-pair order parameters, with respect to two independent sub-systems respectively, are introduced to characterize the phase diagram by its value and nonanalytic behavior at phase boundaries. We find that the ground state are two-fluid condensate, with two different BCS pairing modes. In addition, such a decomposition holds true not only for ground state but also the whole eigen space. Then it should result in the decomposition of dynamics. Based on the exact results on the dynamics obtained in the previous workSYB_PRB , we study the dynamics of the SSH Kitaev model. The primary aim of this study was to obtain pertinent information regarding the nonequilibrium state. In this study, the starting point is the vacuum state, a trivial initial state. We investigate the evolved states under the postquench Hamiltonian with parameters covering the entire region and find two types of order parameters that are different but can help determine the quantum phase diagram. Finally, the robustness of such a decomposition to the perturbation of nonzero chemical potential is also investigated. Analytical analysis and numerical simulations show that the real-space decomposition for the ground state still holds true approximately in many regions in the presence of finite chemical potential.

This paper is organized as follows. In Sec. II, we present a generalized model and introduce a real-space Bogoliubov transformation to decompose the model Hamiltonian. In Secs. III, we apply the transformation on a SSH Kitaev model and deduce the phase diagram from decomposed sub-Hamiltonians. In Secs. IV and V, two types of BCS order parameters are introduced. The explicit expressions for both ground state and non-equilibrium state are obtained, respectively. Sec. VI devotes to the investigation for the case with nonzero potential. Finally, we provide a summary and discussion in Section VII.

Refer to caption
Figure 1: Schematic of phase diagrams for the ground states of systems (a) HAH_{A}, (b) HBH_{B} and (c) HH, respectively. The phase boundaries separating the regions with different colors are the plots of the curves from Eqs. (19), (20), (21), and (22). The winding numbers are denoted in each regions. One can see that all the information in (c) can be obtained directly by that of (a) and (b), demonstrating the benefits of the real-space decomposition.

II Model and decomposition

We consider the following fermionic Hamiltonian on a lattice of length 2N2N

H=j=12N[Jjcjcj+1+Δjcjcj+1+H.c.+μj(2nj1)],H=\sum\limits_{j=1}^{2N}[J_{j}c_{j}^{{\dagger}}c_{j+1}+\Delta_{j}c_{j}^{{\dagger}}c_{j+1}^{{\dagger}}+\mathrm{H.c.}+\mu_{j}\left(2n_{j}-1\right)], (1)

where cjc_{j}^{{\dagger}} (cj)(c_{j}) is a fermionic creation (annihilation) operator on site jj, nj=cjcjn_{j}=c_{j}^{{\dagger}}c_{j}, JjJ_{j} the tunneling rate across the dimer (j,j+1)\left(j,j+1\right), μj\mu_{j} the on-site chemical potential, and Δj\Delta_{j} the strength of the pp-wave pair creation (annihilation) on the dimer (j,j+1)\left(j,j+1\right). c2N+1=c1c_{2N+1}=c_{1} is defined for periodic boundary condition. The Hamiltonian in Eq. (1) has a rich phase diagram that describes a spin-polarized pp-wave superconductor in one dimension when the parameters (Jj,Δj,μj)\left(J_{j},\Delta_{j},\mu_{j}\right) are taken in a uniform fashion Ryohei Wakatsuki . This system has a topological phase in which a zero- energy Majorana mode is located at each end of a long chain by taking the open boundary condition c2N+1=0c_{2N+1}=0. It is the fermionized version of the well-known one-dimensional transverse-field Ising model Pfeuty , which is one of the simplest solvable models when the translational symmetry is imposed that exhibits quantum criticality and phase transition with spontaneous symmetry breaking SachdevBook ; ZG . In addition, several studies have been conducted with a focus on long-range Kitaev chains DV1 ; DV2 ; OV ; LL ; UB . Although system we concerned in this work contains only the nearest neighbor hopping and pairing terms, this method be applied to the system involving long-range terms under certain conditions.

We introduce a linear unitary transformation

(αjαjβjβj)=T(c2j1c2j1c2jc2j)=T(AjAjBjBj),\left(\begin{array}[]{c}\alpha_{j}^{{\dagger}}\\ \alpha_{j}\\ \beta_{j}^{{\dagger}}\\ \beta_{j}\end{array}\right)=T\left(\begin{array}[]{c}c_{2j-1}^{{\dagger}}\\ c_{2j-1}\\ c_{2j}^{{\dagger}}\\ c_{2j}\end{array}\right)=T\left(\begin{array}[]{c}A_{j}^{{\dagger}}\\ A_{j}\\ B_{j}^{{\dagger}}\\ B_{j}\end{array}\right), (2)

with the matrix

T=(T)1=12(11111111iiiiiiii),T=\left(T^{{\dagger}}\right)^{-1}=\frac{1}{2}\left(\begin{array}[]{cccc}1&1&1&-1\\ 1&1&-1&1\\ -i&i&i&i\\ -i&i&-i&-i\end{array}\right), (3)

which ensures that the αj\alpha_{j} and βj\beta_{j} are still fermionic operators, i.e.,

{αj,αj}\displaystyle\left\{\alpha_{j},\alpha_{j^{\prime}}^{{\dagger}}\right\} =\displaystyle= {αj,βj}=δj,j,{αj,βj}=0,\displaystyle\left\{\alpha_{j},\beta_{j^{\prime}}^{{\dagger}}\right\}=\delta_{j,j^{\prime}},\left\{\alpha_{j},\beta_{j^{\prime}}^{{\dagger}}\right\}=0,
{αj,αj}\displaystyle\left\{\alpha_{j},\alpha_{j^{\prime}}\right\} =\displaystyle= {βj,βj}={αj,βj}=0,\displaystyle\left\{\beta_{j},\beta_{j^{\prime}}\right\}=\left\{\alpha_{j},\beta_{j^{\prime}}\right\}=0, (4)

and any transformed Hamiltonian maintains its original physics. It is similar to the Bogoliubov transformation if regarding the parity of the site number (2j2j or 2j12j-1) as spin degree of freedom. Applying the transformation (2), the Hamiltonian HH can be expressed in the form

H=HA+HB+HAB,H=H_{A}+H_{B}+H_{AB}, (5)

where

HA=12j=1N[(J2j+Δ2j)(αjαj+1+αjαj+1)\displaystyle H_{A}=\frac{1}{2}\sum\limits_{j=1}^{N}[\left(J_{2j}+\Delta_{2j}\right)\left(\alpha_{j}^{{\dagger}}\alpha_{j+1}+\alpha_{j}^{{\dagger}}\alpha_{j+1}^{{\dagger}}\right)
+H.c.+(Δ2j1J2j1)(2αjαj1)],\displaystyle+\mathrm{H.c.}+\left(\Delta_{2j-1}-J_{2j-1}\right)\left(2\alpha_{j}^{{\dagger}}\alpha_{j}-1\right)], (6)

and

HB=12j=1N[(Δ2jJ2j)(βjβj+1+βjβj+1)\displaystyle H_{B}=\frac{1}{2}\sum\limits_{j=1}^{N}[\left(\Delta_{2j}-J_{2j}\right)\left(\beta_{j}^{{\dagger}}\beta_{j+1}+\beta_{j}^{{\dagger}}\beta_{j+1}^{{\dagger}}\right)
+H.c.+(J2j1+Δ2j1)(2βjβj1)],\displaystyle+\mathrm{H.c.}+\left(J_{2j-1}+\Delta_{2j-1}\right)\left(2\beta_{j}^{{\dagger}}\beta_{j}-1\right)], (7)

denote two Kitaev chain of length NN and commute with each other

[HA,HB]=0.\left[H_{A},H_{B}\right]=0. (8)

The third term is {μl}\left\{\mu_{l}\right\} dependent, containing the intra- and inter-chain interactions

HAB\displaystyle H_{AB} =\displaystyle= ij=1N[(μ2j1+μ2j)βjαj\displaystyle i\sum_{j=1}^{N}[\left(\mu_{2j-1}+\mu_{2j}\right)\beta_{j}^{{\dagger}}\alpha_{j}^{{\dagger}} (9)
+(μ2j1μ2j)βjαj]+H.c.\displaystyle+\left(\mu_{2j-1}-\mu_{2j}\right)\beta_{j}^{{\dagger}}\alpha_{j}]+\mathrm{H.c.}

Obviously, when consider the case with zero chemical potential, μl=0\mu_{l}=0, the Hamiltonian is exactly decomposed into two independent ones. Accordingly we have

H|ψ=H|ψA|ψB=(εA+εB)|ψ,H\left|\psi\right\rangle=H\left|\psi_{A}\right\rangle\left|\psi_{B}\right\rangle=\left(\varepsilon_{A}+\varepsilon_{B}\right)\left|\psi\right\rangle, (10)

with

HA|ψA=εA|ψA,HB|ψB=εB|ψB.H_{A}\left|\psi_{A}\right\rangle=\varepsilon_{A}\left|\psi_{A}\right\rangle,H_{B}\left|\psi_{B}\right\rangle=\varepsilon_{B}\left|\psi_{B}\right\rangle. (11)

For arbitrary operator functions FA(α1,,F_{A}(\alpha_{1},..., αl,,αN;\alpha_{l},...,\alpha_{N}; α1,,\alpha_{1}^{{\dagger}},..., αj,,αN)\alpha_{j}^{{\dagger}},...,\alpha_{N}^{{\dagger}}) and FB(β1,,F_{B}(\beta_{1},..., βl,,\beta_{l},..., βN;\beta_{N}; β1,,\beta_{1}^{{\dagger}},..., βj,,βN)\beta_{j}^{{\dagger}},...,\beta_{N}^{{\dagger}}), we always have

ψ|FA|ψ\displaystyle\left\langle\psi\right|F_{A}\left|\psi\right\rangle =\displaystyle= ψA|FA|ψA,\displaystyle\left\langle\psi_{A}\right|F_{A}\left|\psi_{A}\right\rangle,
ψ|FB|ψ\displaystyle\left\langle\psi\right|F_{B}\left|\psi\right\rangle =\displaystyle= ψB|FB|ψB.\displaystyle\left\langle\psi_{B}\right|F_{B}\left|\psi_{B}\right\rangle. (12)

Then any features relate to FAF_{A}(FBF_{B}) in chain AA(BB) should emerge in the original system. The the following we will demonstrate this in a concrete example.

III SSH Kitaev model

Now we consider a SSH Kitaev model with the Hamiltonian

H=j=1N[λc2jc2j+1+(1λ)c2j1c2j+λΔc2j1c2j\displaystyle H=\sum_{j=1}^{N}[\lambda c_{2j}^{{\dagger}}c_{2j+1}+\left(1-\lambda\right)c_{2j-1}^{{\dagger}}c_{2j}+\lambda\Delta c_{2j-1}c_{2j}
+(1λ)Δc2jc2j+1+2μ(n2j1+n2j1)]+H.c.,\displaystyle+\left(1-\lambda\right)\Delta c_{2j}c_{2j+1}+2\mu\left(n_{2j-1}+n_{2j}-1\right)]+\mathrm{H.c.}, (13)

in which, both the hopping and pairing strengths are assigned alternatively. It has been well studied in the refs. Ryohei Wakatsuki . In this work, we focus on the connection between this 2N2N-site system and two NN-site sub-systems. When taking μ=0\mu=0, the original Hamiltonian can be written in the form H=HA+HBH=H_{A}+H_{B} with

Hσ\displaystyle H_{\sigma} =\displaystyle= j=1N[Jσ(dj,σdj+1,σ+dj,σdj+1,σ)+H.c.\displaystyle\sum_{j=1}^{N}[J_{\sigma}\left(d_{j,\sigma}^{{\dagger}}d_{j+1,\sigma}^{{\dagger}}+d_{j,\sigma}^{{\dagger}}d_{j+1,\sigma}\right)+\text{H.c.} (14)
+μσ(12dj,σdj,σ),\displaystyle+\mu_{\sigma}\left(1-2d_{j,\sigma}^{{\dagger}}d_{j,\sigma}\right),

where σ=A\sigma=A or BB, dj,A=αjd_{j,A}=\alpha_{j}, dj,B=βjd_{j,B}=\beta_{j}, and the corresponding parameters are

JA\displaystyle J_{A} =\displaystyle= [λ(1λ)Δ]/2,\displaystyle\left[\lambda-\left(1-\lambda\right)\Delta\right]/2, (15)
μA\displaystyle\mu_{A} =\displaystyle= [λΔ+(1λ)]/2,\displaystyle\left[\lambda\Delta+\left(1-\lambda\right)\right]/2, (16)
JB\displaystyle J_{B} =\displaystyle= [λ+(1λ)Δ]/2,\displaystyle-\left[\lambda+\left(1-\lambda\right)\Delta\right]/2, (17)
μB\displaystyle\mu_{B} =\displaystyle= [λΔ(1λ)]/2.\displaystyle\left[\lambda\Delta-\left(1-\lambda\right)\right]/2. (18)

We note that both HAH_{A} and HBH_{B} describe the same system but with different parameters. Although the chemical potential is zero in HH, it is nonzero for sub-Hamiltonians. The phase diagram for each sub-Hamiltonian HσH_{\sigma} is well known, and then can be used to obtain the phase diagram of HH. In fact, the phase boundary of the ground state of HσH_{\sigma} are two points μσ/Jσ=±1\mu_{\sigma}/J_{\sigma}=\pm 1. This maps to two curves

Δ2λ+1=0,\Delta-2\lambda+1=0, (19)

and

2λΔΔ+1=0,2\lambda\Delta-\Delta+1=0, (20)

in the λΔ\lambda\Delta-plane for HAH_{A}, while two curves

Δ+2λ1=0,\Delta+2\lambda-1=0, (21)

and

2λΔΔ1=0,2\lambda\Delta-\Delta-1=0, (22)

for HBH_{B}. According to our above analysis, all the four curves constitute the phase diagram of HH. In Fig. 1, we schematically illustrate the phase diagrams of HAH_{A}HBH_{B} and HH, respectively. We can see that all the information in the ground state of HH, including the gapless lines and winding numbers can be obtained directly by that of HAH_{A} and HBH_{B}, as a demonstration of benefits from the real-space decomposition.

IV BCS order parameters

It has been shown that the pairing order parameter can be utilized to characterize the phase diagram SYB_PRB and the long-range order MES_PRB of the ground state of a simple Kitaev chain. In this section, we focus on the similar investigation in this aspect for the present model based on the real-space decomposition method. We first briefly review the obtained conclusion for the simple Kitaev chain. We take HAH_{A} as an example. We introduce the BCS-pairing operator

O^A,k=i(αkαkαkαk),\widehat{O}_{A,k}=i\left(\alpha_{-k}\alpha_{k}-\alpha_{k}^{{\dagger}}\alpha_{-k}^{{\dagger}}\right), (23)

to characterize pairing channels in k{k} space. Here the Fourier transformation

αk\displaystyle\alpha_{k} =\displaystyle= 1Njeikjαj\displaystyle\frac{1}{\sqrt{N}}\sum\limits_{j}e^{-ikj}\alpha_{j} (24)
=\displaystyle= 12(Ak+Bk+AkBk),\displaystyle\frac{1}{2}\left(A_{k}+B_{k}+A_{-k}^{{\dagger}}-B_{-k}^{{\dagger}}\right),

is applied with

(AkBk)=1Nkeikj(AjBj).\left(\begin{array}[]{c}A_{k}\\ B_{k}\end{array}\right)=\frac{1}{\sqrt{N}}\sum_{k}e^{-ikj}\left(\begin{array}[]{c}A_{j}\\ B_{j}\end{array}\right). (25)

For a given state |ψ\left|\psi\right\rangle, the quantity |ψ|O^A,k|ψ||\left\langle\psi\right|\widehat{O}_{A,k}\left|\psi\right\rangle| measures the rate of transition for a BCS pair at the kk channel. For the ground state |GA\left|\text{{G}}_{A}\right\rangle of HAH_{A} the pairing order parameter of the ground state is expressed as

OA,g=1Nπ>k>0|GA|O^A,k|GA|.O_{A,\mathrm{g}}=\frac{1}{N}\sum_{\pi>k>0}\left|\left\langle\text{{G}}_{A}\right|\widehat{O}_{A,k}\left|\text{{G}}_{A}\right\rangle\right|. (26)

In the large NN limit, it can be expressed explicitly as

OA,g\displaystyle O_{A,\mathrm{g}} =\displaystyle= 1π0πsink2(μA/JAcosk)2+sin2kdk\displaystyle\frac{1}{\pi}\int_{0}^{\pi}\frac{\sin k}{2\sqrt{\left(\mu_{A}/J_{A}-\cos k\right)^{2}+\sin^{2}k}}\mathrm{d}k (29)
=\displaystyle= 1π{1,|μA/JA|1JA/μA,|μA/JA|>1.\displaystyle\frac{1}{\pi}\left\{\begin{array}[]{cc}1,&\left|\mu_{A}/J_{A}\right|\leqslant 1\\ J_{A}/\mu_{A},&\left|\mu_{A}/J_{A}\right|>1\end{array}\right..

Obviously, there exist non-analytic points at |μA/JA|=1\left|\mu_{A}/J_{A}\right|=1, as the signatures of quantum phase boundary. The same results hold for HBH_{B}.

Refer to caption
Figure 2: Color contour plots of numerical results of order parameters (a) OA,gO_{A,\mathrm{g}} and (b) OB,gO_{B,\mathrm{g}} defined in (32) and (33), respectively. (c) and (d) are plots of OA,g+OB,gO_{A,\mathrm{g}}+O_{B,\mathrm{g}} and OA,gOB,gO_{A,\mathrm{g}}-O_{B,\mathrm{g}}, respectively. The system parameters are N=1000N=1000 and J=1J=1. The white dashed lines are a guide to the eye to indicate the phase boundaries presented in Fig. 1. It is clear that the order parameters obtained from subsystems can identify the entire phase diagram.

These conclusions can be directly applied on the ground state |G\left|\text{{G}}\right\rangle of system HH, mapping the non-analytic points |μA/JA|=|μB/JB|=1\left|\mu_{A}/J_{A}\right|=\left|\mu_{B}/J_{B}\right|=1 on the λΔ\lambda\Delta-plane. In fact, introducing two order parameters

OA,g=1Nπ>k>0|G|i(αkαkαkαk)|G|,O_{A,\mathrm{g}}=\frac{1}{N}\sum_{\pi>k>0}\left|\left\langle\text{{G}}\right|i\left(\alpha_{-k}\alpha_{k}-\alpha_{k}^{{\dagger}}\alpha_{-k}^{{\dagger}}\right)\left|\text{{G}}\right\rangle\right|, (30)

and

OB,g=1Nπ>k>0|G|i(βkβkβkβk)|G|,O_{B,\mathrm{g}}=\frac{1}{N}\sum_{\pi>k>0}\left|\left\langle\text{{G}}\right|i\left(\beta_{-k}\beta_{k}-\beta_{k}^{{\dagger}}\beta_{-k}^{{\dagger}}\right)\left|\text{{G}}\right\rangle\right|, (31)

we directly have the explicit expressions of OA,gO_{A,\mathrm{g}} and OB,gO_{B,\mathrm{g}} in the λΔ\lambda\Delta-plane

OA,g=1π{1,regions I, II, and IIIλ(1λ)ΔλΔ+(1λ),otherwise,O_{A,\mathrm{g}}=\frac{1}{\pi}\left\{\begin{array}[]{cc}1,&\text{regions I, II, and III}\\ \frac{\lambda-\left(1-\lambda\right)\Delta}{\lambda\Delta+\left(1-\lambda\right)},&\text{otherwise}\end{array}\right., (32)

where region I is Δ(12λ1,)\Delta\in(-\frac{1}{2\lambda-1},\infty) for 0<λ<0.50<\lambda<0.5, region II is Δ(12λ1,2λ1)\Delta\in(-\frac{1}{2\lambda-1},2\lambda-1) for 0.5<λ<10.5<\lambda<1, and region III is Δ(,2λ1)\Delta\in(-\infty,2\lambda-1) for 0<λ<0.50<\lambda<0.5, while

OB,g=1π{1,regions I, II, and IIIλ+(1λ)ΔλΔ(1λ),otherwise,O_{B,\mathrm{g}}=\frac{1}{\pi}\left\{\begin{array}[]{cc}1,&\text{regions I, II, and III}\\ -\frac{\lambda+\left(1-\lambda\right)\Delta}{\lambda\Delta-\left(1-\lambda\right)},&\text{otherwise}\end{array}\right., (33)

where region I is Δ(,12λ1)\Delta\in(-\infty,\frac{1}{2\lambda-1}) for 0<λ<0.50<\lambda<0.5, region II is Δ(2λ+1,12λ1)\Delta\in(-2\lambda+1,\frac{1}{2\lambda-1}) for 0.5<λ<10.5<\lambda<1, and region III is Δ(2λ+1,)\Delta\in(-2\lambda+1,\infty) for 0<λ<0.50<\lambda<0.5. In the representation of the original Hamiltonian, the physics of the two order parameters are clear, based on the expressions

OA,g+OB,g\displaystyle O_{A,\mathrm{g}}+O_{B,\mathrm{g}} =\displaystyle= 1Nπ>k>0|G|Λk|G|,\displaystyle\frac{1}{N}\sum_{\pi>k>0}\left|\left\langle\text{{G}}\right|\Lambda_{k}\left|\text{{G}}\right\rangle\right|, (34)
OA,gOB,g\displaystyle O_{A,\mathrm{g}}-O_{B,\mathrm{g}} =\displaystyle= 1Nπ>k>0|G|Σk|G|,\displaystyle\frac{1}{N}\sum_{\pi>k>0}\left|\left\langle\text{{G}}\right|\Sigma_{k}\left|\text{{G}}\right\rangle\right|, (35)

with the pairing and current operators

Λk\displaystyle\Lambda_{k} =\displaystyle= i(AkBk+BkAkH.c.),\displaystyle i\left(A_{-k}B_{k}+B_{-k}A_{k}-\text{{H.c.}}\right), (36)
Σk\displaystyle\Sigma_{k} =\displaystyle= i(AkBk+BkAkH.c.).\displaystyle i\left(A_{k}^{{\dagger}}B_{k}+B_{-k}A_{-k}^{{\dagger}}-\text{{H.c.}}\right). (37)

Obviously, OA,g+OB,gO_{A,\mathrm{g}}+O_{B,\mathrm{g}} measures the BCS pair for fermions from two sub-lattices AA and BB, while OA,gOB,gO_{A,\mathrm{g}}-O_{B,\mathrm{g}} the current across two sub-lattices. In Fig. 2, we plot the functions, OA,gO_{A,\mathrm{g}}, OB,gO_{B,\mathrm{g}} and OA,g±OB,gO_{A,\mathrm{g}}\pm O_{B,\mathrm{g}}, respectively. We can see that quantities OA,gO_{A,\mathrm{g}} and OB,gO_{B,\mathrm{g}} reflect partial phase boundary, while both two types of order parameters OA,g±OB,gO_{A,\mathrm{g}}\pm O_{B,\mathrm{g}} can identify the entire phase diagram.

V Non-equilibrium dynamics

Recently, advancements in atomic physics, quantum optics, and nanoscience have allowed the development of artificial systems with high accuracy Jochim ; Greiner . The study of nonequilibrium many-body dynamics presents an alternative approach for accessing a new exotic quantum state with an energy level considerably different from that of the ground state Choi ; Else ; Khemani ; Lindner ; Kaneko ; Tindall ; YXMPRA ; ZXZPRB2 ; TK ; JT1 ; JT2 . Several works show that the quenching dynamics governed by the post-quench Hamiltonian is intimately related to its ground state SYB_PRB ; MH ; LZ . In this section, we turn to the topic of nonequilibrium phenomena in the present Hamiltonian. For a uniform Kitaev chain, it has been shown that the pairing order parameter for a nonequilibrium state obtained through time evolution from an initially prepared vacuum state still help determine the phase diagram SYB_PRB . For the present model, we will investigate this issue based on the obtained results for the sub-systems HAH_{A} and HBH_{B}.

Similarly, we still give a brief review the obtained conclusion for a simple Kitaev chain, such as chain HAH_{A}. It has been shown that the order parameter for a nonequilibrium state is defined as SYB_PRB

O¯A=limT1T0T1Nk|ψ(t)|O^A,k|ψ(t)|dt,\overline{O}_{A}=\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}\frac{1}{N}\sum_{k}\left|\left\langle\psi(t)\right|\widehat{O}_{A,k}\left|\psi(t)\right\rangle\right|\mathrm{d}t, (38)

where the time evolution |ψA(t)=eiHAt|ψA(0)\left|\psi_{A}(t)\right\rangle=e^{-iH_{A}t}\left|\psi_{A}(0)\right\rangle for a particular initial state |ψ(0)\left|\psi(0)\right\rangle, satisfying

αj|ψA(0)=0.\alpha_{j}\left|\psi_{A}(0)\right\rangle=0. (39)

State |ψ(0)\left|\psi(0)\right\rangle is essentially an empty state for a set of fermions {αj}\left\{\alpha_{j}\right\} in real space. In the large NN limit, it can be expressed explicitly as

O¯A\displaystyle\overline{O}_{A} =\displaystyle= 12π0π|(coskμA/JA)sink(coskμA/JA)2+sin2k|dk\displaystyle\frac{1}{2\pi}\int_{0}^{\pi}\left|\frac{\left(\cos k-\mu_{A}/J_{A}\right)\sin k}{\left(\cos k-\mu_{A}/J_{A}\right)^{2}+\sin^{2}k}\right|\mathrm{d}k (42)
=\displaystyle= 12π{1,|μA/JA|1Λ,|μA/JA|>1,\displaystyle\frac{1}{2\pi}\left\{\begin{array}[]{cc}1,&\left|\mu_{A}/J_{A}\right|\leqslant 1\\ \Lambda,&\left|\mu_{A}/J_{A}\right|>1\end{array}\right.,

where

Λ=|JAμA+12(1JA2μA2)ln|μA+JAμAJA||.\Lambda=\left|\frac{J_{A}}{\mu_{A}}+\frac{1}{2}\left(1-\frac{J_{A}^{2}}{\mu_{A}^{2}}\right)\ln\left|\frac{\mu_{A}+J_{A}}{\mu_{A}-J_{A}}\right|\right|. (43)

The non-analytic points at |μA/JA|=1\left|\mu_{A}/J_{A}\right|=1 are the evidently signatures of quantum phase boundary. The same results hold for HBH_{B}.

These conclusions can be directly applied to the dynamics of system HH, mapping the non-analytic points |μA/JA|=|μB/JB|=1\left|\mu_{A}/J_{A}\right|=\left|\mu_{B}/J_{B}\right|=1 on the λΔ\lambda\Delta-plane. Accordingly, the nonequilibrium order parameters are defined for the time evolution |ψ(t)=eiHt|ψ(0)\left|\psi(t)\right\rangle=e^{-iHt}\left|\psi(0)\right\rangle =eiHAteiHBt|ψ(0)=e^{-iH_{A}t}e^{-iH_{B}t}\left|\psi(0)\right\rangle of a particular initial state |ψ(0)\left|\psi(0)\right\rangle. To utilize the conclusion for the two sub-systems, the initial state should be chosen as the vacuum state |ψ(0)=|Vac\left|\psi(0)\right\rangle=\left|\text{{Vac}}\right\rangle for both sets of fermion operators {αj}\left\{\alpha_{j}\right\} and {βj}\left\{\beta_{j}\right\}, i.e.,

αj|Vac=βj|Vac=0,\alpha_{j}\left|\text{{Vac}}\right\rangle=\beta_{j}\left|\text{{Vac}}\right\rangle=0,

rather than the vacuum state of operator {cj}\left\{c_{j}\right\}. The vacuum state can be constructed as the form

|Vac=j=1Nβjαj|0=j=1Ni2(1+c2j1c2j)|0,\left|\text{{Vac}}\right\rangle=\prod_{j=1}^{N}\beta_{j}\alpha_{j}\left|0\right\rangle=\prod_{j=1}^{N}\frac{i}{\sqrt{2}}\left(1+c_{2j-1}^{{\dagger}}c_{2j}^{{\dagger}}\right)\left|0\right\rangle, (44)

which is essentially the ground state of the Hamiltonian HH at the special point with λ=1\lambda=1 and Δ1\Delta\ll-1. Then such a special vacuum state can be prepared in this way in the experiment.

By replacing the parameters {μA/JA,μB/JB}\left\{\mu_{A}/J_{A},\mu_{B}/J_{B}\right\} by {Δ,λ}\left\{\Delta,\lambda\right\}, we obtain the explicit expressions of O¯A\overline{O}_{A} and O¯A\overline{O}_{A} in the λΔ\lambda\Delta-plane

O¯A=12π{1,regions I, II, and IIIΩA,otherwise\overline{O}_{A}=\frac{1}{2\pi}\left\{\begin{array}[]{cc}1,&\text{regions I, II, and III}\\ \Omega_{A},&\text{otherwise}\end{array}\right. (45)

where region I is Δ(12λ1\Delta\in(-\frac{1}{2\lambda-1}, )\infty) for 0<λ<0.50<\lambda<0.5, region II is Δ(12λ1\Delta\in(-\frac{1}{2\lambda-1}, 2λ1)2\lambda-1) for 0.5<λ<10.5<\lambda<1, and region III is Δ(\Delta\in(-\infty, 2λ1)2\lambda-1) for 0<λ<0.50<\lambda<0.5,

O¯B=12π{1,regions I, II, and IIIΩB,otherwise\overline{O}_{B}=\frac{1}{2\pi}\left\{\begin{array}[]{cc}1,&\text{regions I, II, and III}\\ \Omega_{B},&\text{otherwise}\end{array}\right. (46)

where region I is Δ(\Delta\in(-\infty, 12λ1)\frac{1}{2\lambda-1}) for 0<λ<0.50<\lambda<0.5, region II is Δ(2λ+1\Delta\in(-2\lambda+1, 12λ1)\frac{1}{2\lambda-1}) for 0.5<λ<10.5<\lambda<1, and region III is Δ(2λ+1\Delta\in(-2\lambda+1, )\infty) for 0<λ<0.50<\lambda<0.5. Here two factors can be expressed explicitly as

ΩA=|λ(1λ)ΔλΔ+(1λ)+12[1(λ(1λ)ΔλΔ+(1λ))2]ln|2λΔ+1Δ12λ+Δ||,\Omega_{A}=\left|\frac{\lambda-\left(1-\lambda\right)\Delta}{\lambda\Delta+\left(1-\lambda\right)}+\frac{1}{2}\left[1-\left(\frac{\lambda-\left(1-\lambda\right)\Delta}{\lambda\Delta+\left(1-\lambda\right)}\right)^{2}\right]\ln\left|\frac{2\lambda\Delta+1-\Delta}{1-2\lambda+\Delta}\right|\right|, (47)

and

ΩB=|λ+(1λ)ΔλΔ(1λ)+12[1(λ+(1λ)ΔλΔ(1λ))2]ln|2λΔ1Δ1+2λ+Δ||.\Omega_{B}=\left|-\frac{\lambda+\left(1-\lambda\right)\Delta}{\lambda\Delta-\left(1-\lambda\right)}+\frac{1}{2}\left[1-\left(\frac{\lambda+\left(1-\lambda\right)\Delta}{\lambda\Delta-\left(1-\lambda\right)}\right)^{2}\right]\ln\left|\frac{2\lambda\Delta-1-\Delta}{-1+2\lambda+\Delta}\right|\right|. (48)

In Fig. 3, we plot the functions, O¯A\overline{O}_{A}, O¯B\overline{O}_{B} and O¯A±O¯B\overline{O}_{A}\pm\overline{O}_{B}, respectively. We can see that quantities O¯A\overline{O}_{A} and O¯B\overline{O}_{B} reflect partial phase boundary, while both two types of order parameters O¯A±O¯B\overline{O}_{A}\pm\overline{O}_{B} can identify the entire phase diagram. Methodologically, it also demonstrates the benefits of the real-space decomposition.

Refer to caption
Figure 3: Color contour plots of numerical results of dynamic order parameters (a) O¯A\overline{O}_{A} and (b) O¯B\overline{O}_{B} defined in (32) and (33), respectively. (c) and (d) are plots of O¯A+O¯B\overline{O}_{A}+\overline{O}_{B} and O¯AO¯B\overline{O}_{A}-\overline{O}_{B}, respectively. The system parameters are N=1000N=1000 and J=1J=1. The white and black dashed lines are a guide to the eye to indicate the phase boundaries presented in Fig. 1. It is clear that the dynamic order parameters obtained from subsystems can also identify the entire phase diagram.
Refer to caption
Figure 4: The plots of the fidelity of f(λ,Δ)f\left(\lambda,\Delta\right) defined in (61) for several representative values of μ\mu, (a) μ=0\mu=0, (b) μ=0.5\mu=0.5, (c) μ=1.0\mu=1.0, and (d) μ=1.5\mu=1.5. The results are obtained by exact diagonalization for the Hamiltonians with N=100N=100. It shows the ground state remains unchanged within many gapful regions, indicating that the real-space decomposition for the ground state still holds true approximately in the presence of finite μ\mu.

VI Robustness against nonzero chemical potential

The real-space decomposition is exact when the chemical potential is zero. In this section, we investigate the influence of nonzero μ\mu on our above conclusions. The decomposition we proposed is no longer valid, in the presence of nonzero μ\mu. However, we will show that our conclusion for the ground state holds approximately in most of regions for small values of μ\mu. In principle, when small μ\mu switches on, one can treat HABH_{AB} as a perturbation term. The perturbation theory tells us the effect of the perturbation term depends strongly on the eigenstates of HA+HBH_{A}+H_{B}. In the limit case, the energy gaps for the groundstate of both HA+HBH_{A}+H_{B} are sufficiently large comparing to the value of μ\mu, HABH_{AB} should have no effect on the groundstates |GA|GB\left|\text{{G}}_{A}\right\rangle\left|\text{{G}}_{B}\right\rangle of HA+HBH_{A}+H_{B}. On the other hand, at the phase boundary, the gapless point, the ground states |GA|GB\left|\text{{G}}_{A}\right\rangle\left|\text{{G}}_{B}\right\rangle become (or quasi-) degenerate. The term HABH_{AB} may hybridize them and the low-lying excited states, resulting in entangled state, which is deviated from the product state |GA|GB\left|\text{{G}}_{A}\right\rangle\left|\text{{G}}_{B}\right\rangle.

Based on the Fourier transformations for two sub-lattices in Eq. (25), the Hamiltonian with periodic boundary condition can be block diagonalized by this transformation due to its translational symmetry, i.e.,

H=k[0,π]Hk=H0+Hπ+k(0,π)ψkhkψk,H=\sum_{k\in\left[0,\pi\right]}H_{k}=H_{0}+H_{\pi}+\sum_{k\in(0,\pi)}\psi_{k}^{\dagger}h_{k}\psi_{k}, (49)

satisfying [Hk,Hk]=0\left[H_{k},H_{k^{\prime}}\right]=0, where the operator vector ψk=(Ak,Bk,Ak,Bk)\psi_{k}^{\dagger}=\left(A_{k}^{\dagger},B_{k}^{\dagger},A_{-k},B_{-k}\right), and the core matrix is expressed explicitly as

hk=(μz0wzμw00wμzw0zμ),h_{k}=\left(\begin{array}[]{cccc}\mu&z&0&w\\ z^{\ast}&\mu&-w^{\ast}&0\\ 0&-w&-\mu&-z\\ w^{\ast}&0&-z^{\ast}&-\mu\end{array}\right), (50)

where z=λeikz=\lambda e^{ik} +(1λ)+\left(1-\lambda\right) and w=Δ[λ (1λ)eik]w=\Delta\left[\lambda\text{ }-\left(1-\lambda\right)e^{ik}\right]. Here, H0H_{0} and HπH_{\pi} have the form

H0\displaystyle H_{0} =\displaystyle= 2JA0B0+2JB0A0+2μ(A0A0B0B0)\displaystyle 2JA_{0}^{\dagger}B_{0}+2JB_{0}^{\dagger}A_{0}+2\mu\left(A_{0}^{{\dagger}}A_{0}-B_{0}B_{0}^{{\dagger}}\right) (51)
+(ΔaΔb)(A0B0+A0B0),\displaystyle+\left(\Delta_{a}-\Delta_{b}\right)\left(A_{0}^{\dagger}B_{0}^{\dagger}+A_{0}B_{0}\right),
Hπ\displaystyle H_{\pi} =\displaystyle= 2μ(AπAπBπBπ)\displaystyle 2\mu\left(A_{\pi}^{{\dagger}}A_{\pi}-B_{\pi}B_{\pi}^{{\dagger}}\right) (52)
(Δa+Δb)(AπBπ+BπAπ),\displaystyle-\left(\Delta_{a}+\Delta_{b}\right)\left(A_{\pi}^{\dagger}B_{\pi}^{\dagger}+B_{\pi}A_{\pi}\right),

and HπH_{\pi} vanishes when odd NN is taken. To demonstrate the above analysis, one can rewrite the matrix hkh_{k} in the form

hk=hk0+μΓzh_{k}=h_{k}^{0}+\mu\Gamma^{z} (53)

with

hk0=(0z0wz0w00w0zw0z0),h_{k}^{0}=\left(\begin{array}[]{cccc}0&z&0&w\\ z^{\ast}&0&-w^{\ast}&0\\ 0&-w&0&-z\\ w^{\ast}&0&-z^{\ast}&0\end{array}\right), (54)

and a kk-independent matrix

Γz=(1000010000100001).\Gamma^{z}=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&1&0&0\\ 0&0&-1&0\\ 0&0&0&-1\end{array}\right). (55)

The eigen values εσρk\varepsilon_{\sigma\rho}^{k} and vectors ϕσρk\phi_{\sigma\rho}^{k} of hk0h_{k}^{0} can be obtained as

((ϕ++k)T(ϕ+k)T(ϕ+k)T(ϕk)T)=12(ηk1ηk1ηk1ηk1ηk1ηk1ηk1ηk1),\left(\begin{array}[]{c}\left(\phi_{++}^{k}\right)^{T}\\ \left(\phi_{+-}^{k}\right)^{T}\\ \left(\phi_{-+}^{k}\right)^{T}\\ \left(\phi_{--}^{k}\right)^{T}\end{array}\right)=\frac{1}{2}\left(\begin{array}[]{cccc}\eta_{k}&1&-\eta_{k}&1\\ \eta_{k}&-1&\eta_{k}&1\\ \eta_{k}&-1&-\eta_{k}&-1\\ \eta_{k}&1&\eta_{k}&-1\end{array}\right), (56)

and

εσρk=σ|w+ρz|\varepsilon_{\sigma\rho}^{k}=\sigma\left|w+\rho z\right| (57)

satisfying hk0ϕσρk=εσρkϕσρkh_{k}^{0}\phi_{\sigma\rho}^{k}=\varepsilon_{\sigma\rho}^{k}\phi_{\sigma\rho}^{k}, where ηk=w+z|w+z|\eta_{k}=\frac{w+z}{\left|w+z\right|}, with the indices σ,ρ=±\sigma,\rho=\pm. It is easy to check that the hkμh_{k}^{\mu} acts as a flip operator

Γzϕσρk=ϕσ¯ρ¯k,\Gamma^{z}\phi_{\sigma\rho}^{k}=\phi_{\overline{\sigma}\overline{\rho}}^{k}, (58)

with the labels (σ¯,ρ¯)=(σ,ρ)\left(\overline{\sigma},\overline{\rho}\right)=\left(-\sigma,-\rho\right). It directly results in the transition matrix element between energy bands

(ϕσρk)Γzϕσρk=δσσ¯δρρ¯.\left(\phi_{\sigma\rho}^{k}\right)^{{\dagger}}\Gamma^{z}\phi_{\sigma^{\prime}\rho^{\prime}}^{k}=\delta_{\sigma^{\prime}\overline{\sigma}}\delta_{\rho^{\prime}\overline{\rho}}. (59)

This indicates that when the energy gap is sufficiently large compared to the values of the chemical potential,

|μ||εσρkεσ¯ρ¯k|,\left|\mu\right|\ll\left|\varepsilon_{\sigma\rho}^{k}-\varepsilon_{\overline{\sigma}\overline{\rho}}^{k}\right|, (60)

the term μΓz\mu\Gamma^{z} should not affect the eigen vectors of hk0h_{k}^{0}.

We introduce the quantity

f(λ,Δ)=|G(μ)|GA|GB|2=|G(μ)|G(0)|2,f\left(\lambda,\Delta\right)=\left|\langle\text{{G}}(\mu)\left|\text{{G}}_{A}\right\rangle\left|\text{{G}}_{B}\right\rangle\right|^{2}=\left|\langle\text{{G}}(\mu)\left|\text{{G}}(0)\right\rangle\right|^{2}, (61)

to quantitatively measure the effect of μ\mu on the groundstate |G\left|\text{{G}}\right\rangle of HH. Our above analysis predicts that the gapless lines of both HAH_{A} and HBH_{B} result in the valley of f(λ,Δ)f\left(\lambda,\Delta\right). Numerical simulations for f(λ,Δ)f\left(\lambda,\Delta\right) in finite size system can be performed by exact diagonalization. It essentially relates to two ground states of HH with zero and nonzero μ\mu, respectively, while |G(0)=|GA|GB\left|\text{{G}}(0)\right\rangle=\left|\text{{G}}_{A}\right\rangle\left|\text{{G}}_{B}\right\rangle is the ground state for zero μ\mu. For given parameters, f(λ,Δ)f\left(\lambda,\Delta\right) can be obtained by diagonalization of matrix hkh_{k} and the eigenstates of H0+HπH_{0}+H_{\pi}.

Fig. 4 shows the fidelity of f(λ,Δ)f\left(\lambda,\Delta\right) for a given finite system as a function of (λ,Δ)\left(\lambda,\Delta\right) with various values of μ\mu. In the case of zero μ\mu, the fidelity is unitary everywhere as expected. In plot for small μ\mu, there are evidently sudden drops of the value of fidelity at the phase boundaries and the drops become sharper and sharper as μ\mu increases. These behaviors can be ascribed to a dramatic change of the ground state of the system around the phase boundaries. One can also find out that the ground state remains unchanged within many regions. It indicates the real-space decomposition for the ground state still holds true approximately in many regions in the presence of finite μ\mu.

VII Summary

In summary, we have extended the Bogoliubov transformation for spinless fermions in kk-space to the one in real space, which has been shown to be a tool for the spatial decomposition of a class of Kitaev chains. It provides a way to get insight into a complicated system from that of two decoupled sub-systems. A systematic investigation of a SSH Kitaev model is performed, including the ground state property and nonequilibrium behavior of quenching dynamics. In addition, we also studied the approximate decomposition in the case of nonzero chemical potential. Analytical analysis and numerical simulation show that the real-space decomposability is maintained in most regions of the phase diagram when the chemical potential is small. Our findings not only contribute to the methodology for solving the many-body problem, but also reveal the underlying mechanism of the features of the SSH Kitaev chain.

Acknowledgment

We acknowledge the support of NSFC (Grants No. 12374461).

References

  • (1) A. Y. Kitaev, Unpaired Majorana fermions in quantum wires, Phys. Usp. 44, 131 (2001).
  • (2) 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).
  • (3) Wakatsuki R, Ezawa M, Tanaka Y, et al., Fermion fractionalization to Majorana fermions in a dimerized Kitaev superconductor, Phys. Rev. B 90, 014505 (2014).
  • (4) P. Pfeuty, The one-dimensional Ising model with a transverse field, Ann. Phys. (NY) 57, 79 (1970).
  • (5) S. Sachdev, Quantum Phase Transitions (Cambridge University Press, Cambridge, England, 1999).
  • (6) G. Zhang and Z. Song, Topological Characterization of Extended Quantum Ising Models, Phys. Rev. Lett. 115, 177204 (2015).
  • (7) D. Vodola, L. Lepori, E. Ercolessi, A. V. Gorshkov, and G. Pupillo, Kitaev Chains with Long-Range Pairing, Phys. Rev. Lett. 113, 156402 (2014).
  • (8) D. Vodola, L. Lepori, E. Ercolessi, A. V. Gorshkov, and G. Pupillo, Long-range ising and kitaev models: Phases, correlations and edge modes, New. J. Phys. 18, 015001 (2015).
  • (9) O. Viyuela, D. Vodola, G. Pupillo, and M. A. Martin-Delgado, Topological massive dirac edge modes and long-range superconducting hamiltonians, Phys. Rev. B 94, 125121 (2016).
  • (10) L. Lepori and L. Dell’Anna, Long-range topological insulators and weakened bulk-boundary correspondence, New. J. Phys. 19, 103030 (2017).
  • (11) U. Bhattacharya, S.Maity, A. Dutta, and D. Sen, Critical phase boundaries of static and periodically kicked long-range kitaev chain, J. Phys.: Condens. Matter 31, 174003 (2019).
  • (12) E. S. Ma and Z. Song, Off-diagonal long-range order in the ground state of the Kitaev chain, Phys. Rev. B. 107, 205117(2023).
  • (13) S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, S. Riedl, C. Chin, J. Hecker Denschlag, and R. Grimm, Bose-Einstein Condensation of Molecules, Science 302, 2101 (2003).
  • (14) M. Greiner, C. A. Regal, and D. S. Jin, Emergence of a molecular Bose–Einstein condensate from a Fermi gas, Nature (London) 426, 537 (2003).
  • (15) 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).
  • (16) D. V. Else, B. Bauer, and C. Nayak, Floquet time crystals, Phys. Rev. Lett. 117, 090402 (2016).
  • (17) V. Khemani, A. Lazarides, R. Moessner, and S. L. Sondhi, Phase structure of driven quantum systems, Phys. Rev. Lett. 116, 250401 (2016).
  • (18) N. H. Lindner, G. Refael, and V. Galitski, Floquet Topological Insulator in Semiconductor Quantum Wells, Nat. Phys. 7, 490 (2011).
  • (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) 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).
  • (22) X. Z. Zhang and Z. Song, η\eta-pairing ground states in the non-Hermitian Hubbard model, Phys. Rev. B 103, 235153 (2021).
  • (23) T. Kaneko, T. Shirakawa, S. Sorella, and S. Yunoki, Photoinduced η\eta Pairing in the Hubbard Model, Phys. Rev. Lett. 122, 077002 (2019).
  • (24) J. Tindall, F. Schlawin, M. A. Sentef, and D. Jaksch, Analytical solution for the steady states of the driven Hubbard model, Phys. Rev. B 103, 035146
  • (25) J. Tindall, F. Schlawin, M. Sentef and D. Jaksch, Lieb’s Theorem and Maximum Entropy Condensates, Quantum 5, 610 (2021).
  • (26) M. Heyl, Dynamical quantum phase transitions: a review, Rep. Prog. Phys 81, 054001(2018).
  • (27) L. W. Zhou and Q. Q. Du, Non-Hermitian topological phases and dynamical quantum phase transitions: a generic connection, New J. Phys. 23, 063041(2021).