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

Non-thermal dynamics in a spin-1/2 lattice Schwinger model

Chunping Gao1    Zheng Tang1    Fei Zhu1    Yunbo Zhang2 ybzhang@zstu.edu.cn    Han Pu3 hpu@rice.edu    Li Chen1 lchen@sxu.edu.cn 1Institute of Theoretical Physics, State Key Laboratory of Quantum Optics and Quantum Optics Devices, Shanxi University, Taiyuan 030006, China
2Key Laboratory of Optical Field Manipulation of Zhejiang Province and Physics Department of Zhejiang Sci-Tech University, Hangzhou 310018, China
3Department of Physics and Astronomy, and Rice Center for Quantum Materials, Rice University, Houston, TX 77005, USA
Abstract

Local gauge symmetry is intriguing for the study of quantum thermalization breaking. For example, in the high-spin lattice Schwinger model (LSM), the local U(1) gauge symmetry underlies the disorder-free many-body localization (MBL) dynamics of matter fields. This mechanism, however, would not work in a spin-1/2 LSM due to the absence of electric energy in the Hamiltonian. In this paper, we show that the spin-1/2 LSM can also exhibit disorder-free MBL dynamics, as well as entropy prethermalization, by introducing a four-fermion interaction into the system. The interplay between the fermion interaction and U(1) gauge symmetry endows the gauge fields with an effectively disordered potential which is responsible for the thermalization breaking. It induces anomalous (i.e., non-thermal) behaviors in the long-time evolution of such quantities as local observables, entanglement entropy, and correlation functions. Our work offers a new platform to explore emergent non-thermal dynamics in state-of-the-art quantum simulators with gauge symmetries.

I Introduction

Quantum thermalization is prevalent in quantum many-body physics. It refers to the phenomenon that the long-time dynamical behavior of a closed quantum system can be described by a thermal ensemble characterized by a few parameters such as temperature and chemical potential, accompanied by the loss of the local information of the initial state [1, 2, 3, 4]. Two important classes of exceptions are known to severely break quantum thermalization, one is the quantum integrable systems with the number of conserved quantities being equal to the degree of freedoms [5, 6], and the other is the disordered systems which support the many-body localization (MBL) [7, 8, 9]. A strongly disordered system typically carries a set of local integrals of motion which localizes the excitations and freezes the transport [10, 11, 12, 13], allowing the local information of the initial states to survive for a long time without being erased. These features also underlie several potential applications of MBL states in quantum information processing. Over the past decade, MBL has been extensively studied in various contexts of physical systems, including cold atoms in optical lattices [14, 15, 16], trapped ions [17], nuclear magnetic resonance [18], superconducting circuits [19, 20] and so on.

In recent years, another interesting mechanism of non-thermalization has been found in lattice gauge models without disorders, namely the disorder-free MBL [21, 22, 23, 24, 25, 26, 27]. In these systems, the quantum dynamics are constrained by local gauge symmetries, causing a portion of the system effectively to experience a disorder under the gauge-sector average. Particularly for the lattice quantum electrodynamics (QED) [also called the lattice Schwinger model (LSM)] with gauge fields being realized by high spins (S=1S=1) [22], fermions (matter fields) fail to thermalize when relaxed from a clean Néel state. This MBL results from the combined effect of the U(1) gauge symmetry (the Gauss’s Law) and the electric field energy E2E^{2} in the Hamiltonian. However, in the LSM with gauge fields being spin-1/2, a system that has recently been realized in two cold-atom simulators [28, 29, 30], this disorder-free MBL induced by the electric field energy would not occur due to the vanishing of the E2E^{2} term.

In this paper, we show that, contrary to what has been described above, the gauge fields of the spin-1/2 LSM can in fact also exhibit non-thermal dynamics, such as disorder-free MBL and prethermalization, as long as the system carries a four-fermion interaction term. Including such a fermion interaction into the Schwinger model was motivated by a recent proposal on realizing the synthetic U(1) gauge field using spin-1 bosons [31], in which the four-fermion interaction naturally arises from the intrinsic interactions of spinor cold atoms. With the help of Gauss’s Law, the fermion interaction can be transformed away which gives rise to a type of effective disorder for the gauge particles, rendering the latter to exhibit MBL dynamics. We carry out detailed numerical simulations on such quantities as local observables, bipartite entanglement entropy and correlation functions, in which the dynamical features of thermalization breaking can be clearly demonstrated.

The rest of this paper is organized as follows: In Sec. II, we introduce the U(1) lattice Schwinger model and briefly review the mechanism of the disorder-free MBL. In Sec. III, we present our scheme for breaking the thermalization by four-fermion interactions in a spin-1/2 LSM. In Sec. IV, we go into detail about our numerical results. A brief conclusion can be found in Sec. V.

II Dynamical MBL in the High-Spin LSM

Refer to caption
Figure 1: (a) Schematic of the LSM [Eq. (1)] and that with four-fermion interaction [Eq. (3)]. The round circles denote the matter fields, and ovals denote the gauge fields. The blue dashed line labels a building block consisting of two neighboring gauge fields and one matter field in the middle. ω\omega indicates the coupling between matter and the gauge fields; JJ denotes the fermion interaction between two nearest-neighboring matter fields. (b) LSM and its QED picture in the framework composed of particles and anti-particles, i.e., H~\tilde{H} in Eq. (4). (b1) The correspondence between matter (gauge) fields in the left column and the charges (electric fields) in the right column. Specifically, a matter field occupation on odd (even) sites corresponds to the generation of a positron (electron) with a positive (negative) electric charge in the vacuum. An up-polarized gauge spin on odd (even) sites denotes the right-moving (left-moving) electric fields. The electric directions are inverted for a down-polarized gauge spin. (b2) A detailed example of a state (upper row), with all the matter sites being occupied and all the gauge spins being down-polarized, and its QED picture (bottom row).

Before fully engaging in our scheme, we briefly review the disorder-free MBL in the LSM with high-spin gauge fields. The continuous Schwinger model refers to the (1+1)D QED theory with U(1) gauge invariance, depicting the interactions between electrons (matter fermions) and photons (gauge bosons). It is also widely used as a toy model to study various phenomena in quantum chromodynamics, such as quark confinement and chiral symmetry breaking [32, 33, 34]. The lattice Hamiltonian of the Schwinger model can be obtained by following the discretization convention provided by Kogut and Susskind [35], which is formalized as (setting =1\hbar=1)

HLSM=\displaystyle H_{\text{LSM}}= ωj(ψj1Ujψj+h.c.)\displaystyle-\omega\sum_{j}\left(\psi_{j-1}^{\dagger}U_{j}\psi_{j}+\text{h.c.}\right) (1)
+mj(1)jψjψj+g22jEj2,\displaystyle+m\sum_{j}(-1)^{j}\psi_{j}^{\dagger}\psi_{j}+\frac{g^{2}}{2}\sum_{j}E_{j}^{2},

where ψj{\psi}_{j}^{\dagger} (ψj{\psi}_{j}) indicate the local matter fields of charged fermions, j+={1,2,,L}j\in\mathbb{Z}^{+}=\{1,2,...,L\} with LL being the length of the chain; UjU_{j} and EjE_{j} satisfy the 𝔰𝔲\mathfrak{su}(2)-like algebra [Ej,Uk]=δj,kUj[E_{j},U_{k}]=\delta_{j,k}U_{j}, and denote respectively the parallel transporter and the electric field of the gauge fields living on the link between two neighboring matter sites ψj1{\psi}_{j-1} and ψj{\psi}_{j}, as is schematically shown in Fig. 1(a). In HLSMH_{\text{LSM}}, the first term describes the coupling between matter and gauge fields with coupling strength ω\omega, and the second term is the staggered mass referring to the opposite mass experienced by fermions seated on odd and even sites. The occurrence of negative mass is somewhat strange. However, as we will show later in Sec. III, by introducing the antiparticles, the mass term will have a clearer picture — two neighboring fermions respectively correspond to the electron and the positron carrying opposite charges but the same mass. The last term of HLSMH_{\text{LSM}} indicates the energy of the gauge field with g2>0g^{2}>0 the coupling constant, which is purely composed of the electric energy E2E^{2}. This is a property of (1+1)D where the magnetic field is absent since the curl of the vector potential field is forbidden in one-dimensional space. In quantization, the electric states can only take integer values up to a shift, i.e., Ej=θ/2πE_{j}=\mathbb{Z}-\theta/2\pi, where θ[0,2π)\theta\in[0,2\pi) is the topological angle indicating a background electric field [34, 36, 37].

The LSM [Eq. (1)] carries a local gauge symmetry [Gj,HLSM]=0[G_{j},H_{\text{LSM}}]=0, with

Gj=ψjψj(Ej+1Ej)+12[(1)j1],G_{j}=\psi_{j}^{\dagger}\psi_{j}-(E_{j+1}-E_{j})+\frac{1}{2}[(-1)^{j}-1]\,, (2)

being the Gauss operator defined in a building block consisting of two gauge fields {Ej,Ej+1}\{E_{j},E_{j+1}\} and one matter field ψj\psi_{j} in the middle [see Fig. 1(a)]. The static charge qjq_{j} is defined as the quantum number of GjG_{j}, which is apparently a good quantum number. Up to some constants, qjq_{j} locally characterizes the difference between the net electric flux Ej+1EjE_{j+1}-E_{j} and the fermionic charge ψjψj\psi_{j}^{\dagger}\psi_{j}, which is a direct manifestation of the Gauss’s Law. The local gauge symmetry divides the entire Hilbert space into different gauge sectors, with each gauge sector labeled by a set of static charge numbers 𝐪={q1,q2,,qL}\mathbf{q}=\{q_{1},q_{2},...,q_{L}\}.

To quantum simulate the LSM in experiments [28, 29, 30], it is common to realize the electric fields by spin-SS spinors, i.e., Uj()Sj±U^{({\dagger})}_{j}\rightarrow S_{j}^{\pm} and EjSjzE_{j}\rightarrow S_{j}^{z}, which is also called the quantum link model [38, 39] or spin-gauge model [40, 41]. This means selecting a finite-dimensional representation for the 𝔰𝔲\mathfrak{su}(2) gauge fields and truncating EjE_{j} within the range [S,S][-S,S]. For high-spin LSM, the matter fields can exhibit the disorder-free MBL as the gauge fields are integrated out. The electric energy term Ej2E_{j}^{2} in HLSMH_{\text{LSM}} is responsible for this phenomenon. Let us take S=1S=1 as an example [22]. In each gauge sector, the gauge field can be expressed by the static charge qjq_{j} and the matter-field occupation ψjψj\psi^{\dagger}_{j}\psi_{j} using Gauss’s Law [Eq. (2)]. Consequently, the electric energy can be re-expressed completely in terms of the matter fermions, and contains a term Hd=jqjψjψjH_{\text{d}}=\sum_{j}q^{\prime}_{j}\psi^{\dagger}_{j}\psi_{j}, with qj(𝐪)q^{\prime}_{j}(\mathbf{q}) being a function depending on the gauge-sector number 𝐪\mathbf{q}. Therefore, if the initial state |Ψ0\left|\Psi_{0}\right\rangle spans over a large number of random gauge sectors, HdH_{\text{d}} effectively acts like a disorder potential, rendering the post-quench dynamics to break the thermalization. However, this mechanism would no longer work in the spin-1/2 LSM, since in this case, Ej2=(σjz)2/4=1/4E_{j}^{2}=(\sigma_{j}^{z})^{2}/4=1/4 (with σjz\sigma_{j}^{z} being the spin-1/2 Pauli matrix) is simply a constant that can be neglected. Hence, the electric-energy-based approach in inducing the disorder-free MBL will not be applicable.

We additionally would like to mention that in the gauge sector 𝐪=𝟎\mathbf{q}=\mathbf{0}, the E2E^{2} term is also closely related to the charge confinement of QED [42, 41, 34]. Particularly, the high-spin LSM (θπ\theta\neq\pi) is confined. Separating two fermions with opposite charges would lead to an electric string between them, and hence the total electric energy in the Hamiltonian would be linearly proportional to the length of the string, i.e., |ij|g2S2\propto|i-j|g^{2}S^{2}. To lower the electric energy, an additional pair of fermionic charges will emerge to screen the electric string, which is known as the string breaking [42, 41, 43]. In contrast, the LSM with S=1/2S=1/2 (θ=π\theta=\pi) is deconfined due to the absence of the electric energy in HLSMH_{\text{LSM}}. In this case, the state of a local electric field should be either 1/21/2 or 1/2-1/2, causing the total electric energy to be a constant Lg2S2/2Lg^{2}S^{2}/2 independent of the distribution of fermions. Note that not all gauge sectors are equivalent for the LSM with a finite SS, and hence the relationship between the confinement and the disorder-free localization remains an open question worthy of further study. However, this question goes beyond the scope of our current paper.

III Dynamical MBL in the Spin-1/2 LSM

Considering the fact that current state-of-the-art experimental techniques can only realize the spin-1/2 LSM [28, 29, 30], it is highly desirable to investigate possible ways to break the thermalization in such a system. Here, we formally discuss our scheme. The basic idea is to introduce a four-fermion interaction term into HLSMH_{\text{LSM}} such that the total Hamiltonian now reads,

H=\displaystyle H= ωj(ψj1Sj+ψj+h.c.)\displaystyle-\omega\sum_{j}\left(\psi_{j-1}^{\dagger}S_{j}^{+}\psi_{j}+\text{h.c.}\right) (3)
+mj(1)jψjψj+Jjψj1ψj1ψjψj,\displaystyle+m\sum_{j}(-1)^{j}\psi_{j}^{\dagger}\psi_{j}+J\sum_{j}{\psi}_{j-1}^{\dagger}{\psi}_{j-1}{\psi}_{j}^{\dagger}{\psi}_{j},

with JJ characterizing the interaction strength. The addition of this interaction term does not affect the local gauge symmetry as [Gj,H]=0[G_{j},H]=0 still holds. Our goal is to show that, by integrating the matter field out, the gauge field experiences an effective disorder. Our approach thus is in contrast to the scheme shown in Ref. [22] where the MBL is realized on matter fields. Introducing the fermion interaction term into the LSM was proposed in Ref. [31], in which the equilibrium-state phase diagram and quench dynamics were studied under a fixed gauge sector 𝐪=𝟎\mathbf{q}=\mathbf{0}. In the current work, we find that the fermion interaction is capable of inducing non-thermal dynamics when different gauge sectors are mixed.

We explicitly introduce the anti-particles by taking the particle-hole transformation on the odd sites, i.e., ψjoddψjodd\psi_{j\in\text{odd}}\rightarrow\psi_{j\in\text{odd}}^{\dagger}, and making a similar transformation on the gauge fields Sjodd+Sjodd,SjoddzSjoddzS_{j\in\text{odd}}^{+}\rightarrow-S_{j\in\text{odd}}^{-},S_{j\in\text{odd}}^{z}\rightarrow-S_{j\in\text{odd}}^{z}, which transforms Hamiltonian (3) into a new form:

H~=\displaystyle\tilde{H}= ωj(ψj1Sj+ψj+h.c.)\displaystyle-\omega\sum_{j}\left(\psi_{j-1}S_{j}^{+}\psi_{j}+\text{h.c.}\right) (4)
+mjψjψjJjψj1ψj1ψjψj,\displaystyle+m^{\prime}\sum_{j}\psi_{j}^{\dagger}\psi_{j}-J\sum_{j}\psi_{j-1}^{\dagger}\psi_{j-1}\psi_{j}^{\dagger}\psi_{j},

with m=m+Jm^{\prime}=m+J. Correspondingly, we have the Gauss operator

G~j=ψjψj+Sjz+Sj+1z,\tilde{G}_{j}=\psi_{j}^{\dagger}\psi_{j}+{S}_{j}^{z}+{S}_{j+1}^{z},\\ (5)

with [G~j,H~]=0[\tilde{G}_{j},\tilde{H}]=0, and 𝐪~={q~1,q~2,,q~L}\mathbf{\tilde{q}}=\{\tilde{q}_{1},\tilde{q}_{2},...,\tilde{q}_{L}\} labels the gauge sectors with q~j\tilde{q}_{j} being the quantum number of G~j\tilde{G}_{j}. Apparently, H~\tilde{H} is translationally invariant with all fermions featuring the same mass mm^{\prime}, as mentioned before. H~\tilde{H} provides a clear analog of the LSM in QED [see Fig. 1(b1)]: The occupation of the odd and even matter sites respectively denote the positron and electron with equal mass mm^{\prime}; for gauge spins at even sites, states |\left|\uparrow\right\rangle and |\left|\downarrow\right\rangle respectively correspond to the left- and right-pointing electric fields; whereas for gauge spins at odd sites, the directions of electric fields are reversed. In Fig. 1(b2), we show a concrete example of state (upper row) with all the matter sites being occupied and its QED analog (bottom row), in which the distributions of charges and electric fields are clearly illustrated. In this picture, the matter-gauge interaction (ω\omega term in H~\tilde{H}) indicates the process that a pair of electron and positron merge together simultaneously generating gauge photons. Photon generation in the context of S=1/2S=1/2 corresponds to the spin flip of gauge spins. Also within this picture, Gauss’s Law with G~j\tilde{G}_{j} indicates that the total excitation within a building block is conserved, including the electron (positron) and gauge spins.

Since matter fields and gauge spins are mutually related to each other by Gauss’s Law [Eq. (5)], we are in principle allowed to eliminate the matter fields and write down an effective model purely in terms of the gauge spins. Eliminating the matter fields is straightforward for the last two terms of H~\tilde{H}. To be specific, given a certain gauge sector 𝐪~\mathbf{\tilde{q}}, substituting Eq. (5) into Eq. (4) leads to 2mjSjzJj[q~j1(Sj1z+Sjz)][q~j(Sjz+Sj+1z)]-2m^{\prime}\sum_{j}S_{j}^{z}-J\sum_{j}[\tilde{q}_{j-1}-({S}_{j-1}^{z}+{S}_{j}^{z})][\tilde{q}_{j}-({S}_{j}^{z}+{S}_{j+1}^{z})]. The mm^{\prime}-term is free of disorder, and thus is irrelevant to the MBL dynamics. In the following discussion, we thus focus on the case of m=0m^{\prime}=0. In contrast, the JJ-term, arising from the fermion interaction, is gauge-sector relevant. Rewriting the JJ-term in terms of gauge spins yields

Jj(2SjzSj+1z+Sj1zSj+1zq~jSjz),-J\sum_{j}(2S_{j}^{z}S_{j+1}^{z}+S_{j-1}^{z}S_{j+1}^{z}-\tilde{q}_{j}^{\prime}S_{j}^{z}), (6)

with q~j=q~j2+q~j1+q~j+q~j+1\tilde{q}_{j}^{\prime}=\tilde{q}_{j-2}+\tilde{q}_{j-1}+\tilde{q}_{j}+\tilde{q}_{j+1}. It indicates that, in addition to the homogeneous interactions (SjzSj+1zS_{j}^{z}S_{j+1}^{z} and Sj1zSj+1zS_{j-1}^{z}S_{j+1}^{z}), the gauge field additionally experiences a local potential q~jSjz-\tilde{q}_{j}^{\prime}S_{j}^{z} whose strength depends on the gauge sector 𝐪~\mathbf{\tilde{q}}. Therefore, if the initial state mixes various random gauge sectors, the gauge spins would experience an effective disorder under sector average. This term therefore plays a central role in our scheme in inducing the anomalous non-thermal dynamics, as will be presented in Sec. IV.

q~j\tilde{q}_{j} -1 0 1 2
configurations
|Sjz,nj,Sj+1z\left|S_{j}^{z},n_{j},S_{j+1}^{z}\right\rangle
|,0,\left|\downarrow,0,\downarrow\right\rangle
|,0,\left|\uparrow,0,\downarrow\right\rangle
|,0,\left|\downarrow,0,\uparrow\right\rangle
|,1,\left|\downarrow,1,\downarrow\right\rangle
|,1,\left|\uparrow,1,\downarrow\right\rangle
|,1,\left|\downarrow,1,\uparrow\right\rangle
|,0,\left|\uparrow,0,\uparrow\right\rangle
|,1,\left|\uparrow,1,\uparrow\right\rangle
Table 1: Allowed configurations |Sjz,nj,Sj+1z\left|S_{j}^{z},n_{j},S_{j+1}^{z}\right\rangle in the jj-th building block, with q~j\tilde{q}_{j} being the quantum number of G~j\tilde{G}_{j}.

Within a building block as defined in Fig. 1(a), q~j\tilde{q}_{j} is allowed to take four integer values, i.e., q~j{1,0,1,2}\tilde{q}_{j}\in\{-1,0,1,2\}. By respectively choosing the Fock basis |nj=0,1|n_{j}=0,1\rangle and the spin basis |Siz=,|S_{i}^{z}=\uparrow,\downarrow\rangle for the matter and gauge fields, the correspondence between q~j\tilde{q}_{j} and the allowed configurations is listed in the Tab. 1. It can be observed that q~j={0,1}\tilde{q}_{j}=\{0,1\} each possesses three distinct configurations, whereas q~j={1,2}\tilde{q}_{j}=\{-1,2\} each possesses only one unique configuration. We thus consider an initial state

|Ψ0=(|0+|12)L|,,,,|\Psi_{0}\rangle=\left(\frac{|0\rangle+|1\rangle}{\sqrt{2}}\right)^{\otimes L}|\downarrow,\uparrow,\downarrow,...\rangle, (7)

which is a product state, with the matter fields being an equal superposition of states |0|0\rangle and |1|1\rangle, and the gauge fields being simply an antiferromagnetic Néel state. In each building block, the state |Ψ0|\Psi_{0}\rangle completely lies in q~j={0,1}\tilde{q}_{j}=\{0,1\} with equal probability 1/21/2. Hence, for a chain with length LL, there are totally 2L2^{L} gauge sectors involved. Most of these gauge sectors are with a random 𝐪~\mathbf{\tilde{q}}, e.g., 𝐪~={1,0,0,1,0,1,}\mathbf{\tilde{q}}=\{1,0,0,1,0,1,\cdots\}. There are indeed some exceptions. For example, 𝐪~=𝟎\mathbf{\tilde{q}}=\mathbf{0} and 𝐪~=𝟏\mathbf{\tilde{q}}=\mathbf{1} are completely ordered. However, the portion of them is always exponentially small, and hence they would not dominate the dynamics for a large LL.

Eliminating the matter fields from the first term of H~\tilde{H} is not as straightforward as from the last two terms. Up to now, no simple way exists to eliminate the matter fields for a general random 𝐪~\mathbf{\tilde{q}}. However, as will be shown by numerics below, the ω\omega term alone in Eq. (4) is unable to prevent thermalization, manifested by the phenomenon that the local gauge spins of |Ψ0|\Psi_{0}\rangle quickly relax to thermal equilibrium. Therefore, |Ψ0|\Psi_{0}\rangle serves as an important reference state for the discussion of the thermalization breaking induced by the fermion interaction JJ. It may also be worthwhile to mention that, in the completely ordered gauge sectors (𝐪~=𝟎\mathbf{\tilde{q}}=\mathbf{0} or 𝐪~=𝟏\mathbf{\tilde{q}}=\mathbf{1}), matter-field elimination can be accomplished by mapping the system into a Rydberg chain [44, 31, 36, 45]. The resulting term is a PXP Hamiltonian which is known to possess a set of quantum many-body scar states weakly breaking the eigenstate thermalization hypothesis [46, 47]. In spite of this, the mapping cannot be simply generalized to a general 𝐪~\mathbf{\tilde{q}}. Since the weight of the ordered sectors is sufficiently small as mentioned above, we will not discuss this any further in this paper.

IV Numerical Results

In practical simulations, it is convenient for us to additionally map the fermions of Eq. (4) into Pauli spins using the Jordan-Wigner transformation:

ψj=sj+l=1j1(2nl1),ψj=sjl=1j1(2nl1),\psi^{\dagger}_{j}=s^{+}_{j}\,\prod_{l=1}^{j-1}(2n_{l}-1)\,,\;\;\;\psi_{j}=s^{-}_{j}\,\prod_{l=1}^{j-1}(2n_{l}-1)\,,

with nl=sl+sln_{l}=s^{+}_{l}s^{-}_{l}. Under this mapping, H~\tilde{H} can be written into an interacting spin chain Hamiltonian

Hs=j[ω(sj1Sj+sj+h.c.)+Jsj1zsjz+Jsjz],H_{s}=-\sum_{j}\left[\omega\left(s^{-}_{j-1}S_{j}^{+}s^{-}_{j}+\text{h.c.}\right)+Js^{z}_{j-1}s^{z}_{j}+Js^{z}_{j}\right], (8)

in which the gauge spins and the matter spins are denoted by capital SjS_{j} and lowercase sjs_{j}, respectively. In correspondence, the initial state has the form

|Ψ0=(|+|2)L|,,,,|\Psi_{0}\rangle=\left(\frac{\left|\Uparrow\right\rangle+\left|\Downarrow\right\rangle}{\sqrt{2}}\right)^{\otimes L}|\downarrow,\uparrow,\downarrow,...\rangle, (9)

with |\left|\Uparrow\right\rangle and |\left|\Downarrow\right\rangle denoting the eigenstates of matter spins szs^{z}. We simulate the dynamics |Ψ(t)=eiHst|Ψ0|\Psi(t)\rangle=e^{-iH_{s}t}|\Psi_{0}\rangle via exact diagonalization of the Hamiltonian HsH_{s}. By utilizing the (discrete) translational symmetry of HsH_{s} and |Ψ0|\Psi_{0}\rangle [48, 49], we are able to deal with a system of size up to L=14L=14 (i.e., 14 matter spins plus 14 gauge spins) on a medium-size workstation.

We first look at the dynamics of local polarization of gauge spins, i.e., Sjz(t)\langle S_{j}^{z}(t)\rangle. Generally for a many-body system under thermalization [50, 51, 52], after a sufficiently long time of evolution, all the local information of the initial state would be erased and the system would behave like a thermal state characterized by density matrix ρth\rho_{\text{th}}. Namely, the local observable Sjz(t)\langle S_{j}^{z}(t)\rangle would approach the thermal equilibrium, i.e.,

limtSjz(t)Sjzth=Tr(ρthSjz),\lim_{t\rightarrow\infty}\langle S_{j}^{z}(t)\rangle\approx\langle S_{j}^{z}\rangle_{\text{th}}=\mathrm{Tr}(\rho_{\text{th}}S_{j}^{z})\,, (10)

with

ρth=eβHsTr(eβHs)\rho_{\text{th}}=\frac{e^{-\beta H_{s}}}{\text{Tr}(e^{-\beta H_{s}})} (11)

being the density matrix of the Gibbs ensemble, with β\beta the effective inverse temperature determined by the initial state via =Ψ0|Hs|Ψ0=Tr(ρthHs)\mathcal{E}={\langle}\Psi_{0}|H_{s}|\Psi_{0}\rangle=\text{Tr}(\rho_{\text{th}}H_{s}). In contrast, for systems breaking the thermalization, such as the MBL, the local equilibration limtSjz(t)\lim_{t\rightarrow\infty}\langle S_{j}^{z}(t)\rangle would deviate from the thermal value Sjzth\langle S_{j}^{z}\rangle_{\text{th}}. Our numerics show that, for arbitrary JJ, the thermal state ρth\rho_{\text{th}} associated with our initial state |Ψ0|\Psi_{0}\rangle [Eq. (9)] is always an infinite-temperature thermal state, i.e., ρth𝕀\rho_{\text{th}}\propto\mathbb{I}, such that Sjzth=0\langle S_{j}^{z}\rangle_{\text{th}}=0. This can be understood in the following way. Since |Ψ0|\Psi_{0}\rangle is a product state with each matter spin being (|+|)/2(\left|\Uparrow\right\rangle+\left|\Downarrow\right\rangle)/\sqrt{2} and each gauge spin being either |\left|\uparrow\right\rangle or |\left|\downarrow\right\rangle, it thus has zero energy expectation =Ψ0|Hs|Ψ0=0\mathcal{E}={\left\langle\Psi_{0}\right|H_{s}\left|\Psi_{0}\right\rangle}=0. On the other hand, HsH_{s} is traceless such that the average of all the eigen energies is also equal to zero. These two facts indicate that =Tr(ρthHs)=0\mathcal{E}=\mathrm{Tr}(\rho_{\text{th}}H_{s})=0 should occur at β=0\beta=0, namely at the infinite temperature. The infinite-temperature state should have vanishing expectation values for all the traceless operators, and hence the deviation of the long-time dynamics of local traceless operators from zero conveniently measures the degree of thermalization breaking.

Refer to caption
Figure 2: (a) Time evolution of a local gauge spin Sj=2z\left\langle S^{z}_{j=2}\right\rangle (a1) and a local matter spin sj=2x\left\langle s^{x}_{j=2}\right\rangle (a2) on log(t)\log(t), with solid, dashed, dot-dashed and dotted lines corresponding to the cases of J=0J=0, ω\omega, 2ω2\omega and 3ω3\omega, respectively. The (a3) indicates the averaged long-time polarization S2z¯\overline{S^{z}_{2}} [Eq. (12)] versus system size, where markers are numerical data at L={8,10,12,14}L=\{8,10,12,14\} and lines denote the linear fitting for the data. (b) Dynamics of local gauge polarizations Sjz\left\langle S^{z}_{j}\right\rangle on each site jj, where (b1), (b2) and (b3) denote the cases of J=0J=0, ω\omega and 2ω2\omega, respectively. Except for (a3), all other panels are calculated at L=12L=12.

In Fig. 2(a1), we plot the polarization of a local gauge field Sj=2z(t)\langle S_{j=2}^{z}(t)\rangle for various matter-field interaction JJ, with the solid, dashed, dot-dashed, and dotted lines denoting J=0J=0, ω\omega, 2ω2\omega, and 3ω3\omega, respectively. One can observe that, in the absence of fermion interaction (J=0J=0), the local polarization rapidly decays from 1/21/2 to Sjzth=0\langle S^{z}_{j}\rangle_{\text{th}}=0, as a manifestation of quantum thermalization. However, for the cases of J0J\neq 0, the long-time behaviors Sj=2z(t)\langle S_{j=2}^{z}(t)\rangle apparently deviate from zero. With an increase in JJ, the deviation would become larger. These behaviors are consistent with our previous discussion that increasing JJ leads to an increase in the disorder strength, which results in more severe destruction of quantum thermalization.

By contrast, the matter field does not exhibit thermalization breaking. This is show in Fig. 2(a2), where we plot the dynamics of sj=2x(t)\langle s_{j=2}^{x}(t)\rangle for various values of JJ. As one can see, in the long-time limit, sj=2x(t)\langle s_{j=2}^{x}(t)\rangle all converges to the thermal equilibrium value which is also zero, regardless of the values of JJ. This is understandable since the matter fields do not experience the disorder potential, which thus differs from the gauge fields.

Furthermore, to characterize the dependence of local polarization on the system size LL, we perform a system-size analysis of the averaged polarization S2z¯\overline{S^{z}_{2}} and show the result in the Fig. 2(a3), where

S2z¯=T1t0t0+T𝑑tS2z(t)\overline{S^{z}_{2}}=T^{-1}\int_{t_{0}}^{t_{0}+T}{dt\langle S^{z}_{2}(t)\rangle} (12)

with t0=50ω1t_{0}=50\omega^{-1} and T=300ω1T=300\omega^{-1} chosen to be sufficiently large to ensure that S2z¯\overline{S^{z}_{2}} can capture the averaged long-time feature of the local gauge spin. It can be observed that, for large JJ (J=2ωJ=2\omega and 3ω3\omega), the local polarization increases slowly with system size, which indicates that the system is not ergodic in the thermodynamic limit.

The thermalization process is generally accompanied by the information loss of the initial state, which can be observed in the dynamics of gauge spins as shown in Fig. 2(b1)-(b3). We show the dynamics of Sjz(t)\langle S_{j}^{z}(t)\rangle for each site jj, with panels (b1), (b2) and (b2) corresponding to J=0J=0, ω\omega, and 2ω2\omega, respectively. At t=0t=0, the staggered magnetization for the initial Néel state of gauge spins [Eq. (9)] is quite obvious. As time passes, the staggered magnetization structure vanishes for the case of J=0J=0 [Fig. 2(b1)], indicating the information loss of the initial state. In contrast, for the non-thermal dynamics with J=ωJ=\omega and 2ω2\omega [Fig. 2(b2), (b3)], the staggered magnetization structure persists after a long time of evolution. Moreover, the larger JJ is, the more information about the initial state has remained.

Refer to caption
Figure 3: Bipartite entropy dynamics versus log(t)\log(t), where the columns (a) and (b) respectively correspond to two ways (I and II) of partition of the system. Panels (a1) and (b1) correspond to the von Neumann entropy SI1(t)S^{1}_{\text{I}}(t) and SII1(t)S^{1}_{\text{II}}(t), where the solid, dashed, dot-dashed and dotted lines indicate the cases of J=0J=0, ω\omega, 2ω2\omega and 3ω3\omega, respectively. Panels (a2) and (b2) correspond to the 2nd-order Rényi entropy SI2(t)S^{2}_{\text{I}}(t) and SII2(t)S^{2}_{\text{II}}(t). Panels (a3) and (b3) respectively show the dependence of SI1(t)/LS^{1}_{\text{I}}(t)/L and SII1(t)/LS^{1}_{\text{II}}(t)/L on various lattice sizes LL at a fixed J=3ωJ=3\omega, with solid, dashed, dot-dashed and dotted lines denoting the cases of L=8L=8, 1010, 1212 and 1414.

To characterize the entropy growth in the system, we calculate the dynamics of Rényi entropy

𝒮I,IIα(t)=11αlogTrρAα(t),{\cal S}_{\text{I,II}}^{\alpha}(t)=\frac{1}{1-\alpha}\log\mathrm{Tr}\rho_{\text{A}}^{\alpha}(t), (13)

where ρA=TrBρ=TrB|ΨΨ|\rho_{\text{A}}=\mathrm{Tr}_{\text{B}}\rho=\mathrm{Tr}_{\text{B}}|\Psi\rangle\langle\Psi| is the reduced density matrix of the subsystem A, and α\alpha is the order of Rényi entropy. Particularly in the limit of α1\alpha\rightarrow 1, the Rényi entropy reproduces the von Neumann entropy [53], i.e., 𝒮I,II1=Tr(ρAlogρA){\cal S}^{1}_{\text{I,II}}=-\mathrm{Tr}(\rho_{\text{A}}\log\rho_{\text{A}}). The subscription I and II indicate two different ways of partition the system: I) A consists of left half of gauge spins, while B the rest (i.e., right half of gauge spins and all matter spins); II) A consists of the left half of system including both gauge and matter spins, while B the right half of the system. In partition I, the boundary between the two subsystems is extensive, while in partition II, the boundary is not extensive since it is just a single site as the entire chain is cut into two halves directly from the middle.

In the upper two rows of Fig. 3, we fix L=12L=12 and show respectively the dependence of the von Neumann entropy SI,II1S^{1}_{\text{I,II}} and the 2nd-order Rényi entropy 𝒮I,II2{\cal S}^{2}_{\text{I,II}} on log(t)\log(t), where different line styles again indicate the cases of different fermion interaction JJ. Clearly, for a given partition (I or II), 𝒮1\mathcal{S}^{1} and 𝒮2\mathcal{S}^{2} exhibit similar behavior, allowing us to focus solely on the first row. SI1S^{1}_{\text{I}} and SII1S^{1}_{\text{II}} exhibit a similar long-time behavior after equilibration, i.e., the entropy saturates at a value 𝒮sat{\cal S}_{\text{sat}}. 𝒮sat{\cal S}_{\text{sat}} decreases as JJ increases, akin to the results observed in the conventional disorder-free MBL [22, 25, 26]. However, in the short-time scale, SI,II1S^{1}_{\text{I,II}} exhibit some unconventional features. Particularly in Fig. 3(a1), for J=0J=0, SI1S^{1}_{\text{I}} shows a smooth and rapid growth with speed faster than log(t)\log(t); whereas for large JJ (e.g., J=3ωJ=3\omega), SI1S^{1}_{\text{I}} first hits a small plateau 𝒮pre\mathcal{S}_{\text{pre}}, and then increases approximately linearly in log(t)\log(t) until saturation. The small plateau 𝒮pre\mathcal{S}_{\text{pre}} is called the prethermalization [54, 1, 2, 3] indicating the gauge spins exhibit an intermediate quasi-stationary state before being further thermalized. The prethermalization plateau becomes more and more obvious as JJ grows. SII1S^{1}_{\text{II}} in Fig. 3(a2) is similar to SI1S^{1}_{\text{I}} in the short time, but there exists the difference which mainly lies in that SII1S^{1}_{\text{II}} oscillates during the prethermalization stage of SI1S^{1}_{\text{I}}.

In Fig. 3(a3) and (b3), we fix J=3ωJ=3\omega and show respectively the dependence of 𝒮I1(t)/L{\cal S}^{1}_{\text{I}}(t)/L and 𝒮II1(t)/L{\cal S}^{1}_{\text{II}}(t)/L on various system sizes LL, with solid, dashed, dot-dashed and dotted lines correspond to the cases of L=8L=8, 1010, 1212 and 1414, respectively. The long-time feature of the two figures is quite similar that all curves roughly collapse into a single curve, indicating extensive entropy saturation. On the other hand, the short-time behavior of Fig. 3(a3) is also extensive, whereas that of (b3) is non-extensive. The discrepancy can be attributed to the ways on how the system is partitioned. As we mentioned earlier, the boundary between the two subsystems is extensive (non-extensive) for partition I (II).

Refer to caption
Figure 4: Dynamics of the connected correlation function Γr\Gamma_{r} of gauge spins, with rr being the distance between two spins. (a) The case of J=0J=0. (b) The case of J=3ωJ=3\omega. In the calculation, we fix L=14L=14.

The magnitude of thermalization can be also reflected in the propagation of correlators. In practice, we calculate the connected two-point correlation function of gauge spins:

Γr(t)=Sjz(t)Sj+rz(t)Sjz(t)Sj+rz(t),\displaystyle\Gamma_{r}(t)=\left\langle S^{z}_{j}(t)S^{z}_{j+r}(t)\right\rangle-\left\langle S^{z}_{j}(t)\right\rangle\left\langle S^{z}_{j+r}(t)\right\rangle, (14)

with rr denoting the relative distance. The results of the cases J=0J=0 and J=3ωJ=3\omega are shown in Figs. 4(a) and (b), respectively. One can observe that Γr\Gamma_{r} is zero at t=0t=0 since the initial state |Ψ0|\Psi_{0}\rangle is a product state and also an eigenstate of SjzS^{z}_{j}. As tt increases, Γr\Gamma_{r} spreads out from the center to both sides.

One apparent feature is that the correlation propagation of J=3ωJ=3\omega is much slower than that of J=0J=0, which is consistent with our expectation on MBL [1]. Generally, for a thermalizing system, correlation propagates ballistically forming a light cone |r|t|r|\sim t. In contrast, due to the exponential decay of interaction strength of localized dressed spins, the light core of the MBL is generally in the shape of |r|log(t)|r|\sim\log(t). The correlation boundary in Figs. 4(a) and (b) qualitatively capture the ballistic and log(t)\log(t) light cones, respectively.

V Conclusion

We have shown that the four-fermion interaction term in the spin-1/2 lattice Schwinger model is responsible for the breaking of quantum thermalization. Under the gauge sector average, the gauge spins effectively experience a disorder after the matter degree of freedom is integrated out. This fermion-interaction-induced disorder underlies such non-thermal dynamics as many-body localization and entropy prethermalization when the system relaxes from an antiferromagnetic state. Our work promisingly facilitates the observation of disorder-free many-body localization in state-of-the-art cold-atom quantum simulators with U(1) gauge invariance.

Acknowledgements.
L. C. acknowledges support from the NSF of China (Grants Nos. 12174236 and 12147215); Y. Z. acknowledges support from the NSF of China (Grant No. 12074340) and the Science Foundation of Zhejiang Sci-Tech University (Grant No. 20062098-Y.); H. P. acknowledges support from the US NSF and the Welch Foundation (Grant No. C-1669).

References

  • [1] R. Nandkishore and D. A. Huse, Many-Body Localization and Thermalization in Quantum Statistical Mechanics, Annu. Rev. Condens. Matter Phys. 6, 15 (2015).
  • [2] T. Mori, T. N. Ikeda, E. Kaminishi, and M. Ueda, Thermalization and Prethermalization in Isolated Quantum Systems: A Theoretical Overview, J. Phys. B: At. Mol. Opt. Phys. 51, 112001 (2018).
  • [3] M. Ueda, Quantum Equilibration, Thermalization and Prethermalization in Ultracold Atoms, Nat. Rev. Phys. 2, 669 (2020).
  • [4] D. A. Abanin, E. Altman, I. Bloch, and M. Serbyn, Colloquium: Many-Body Localization, Thermalization, and Entanglement, Rev. Mod. Phys. 91, 021001 (2019).
  • [5] B. Sutherland, Beautiful Models: 70 Years of Exactly Solved Quantum Many-body Problems (World Scientific, River Edge, NJ, 2004).
  • [6] M. Rigol, V. Dunjko, V. Yurovsky, and M. Olshanii, Relaxation in a Completely Integrable Many-Body Quantum System: An Ab Initio Study of the Dynamics of the Highly Excited States of 1D Lattice Hard-Core Bosons, Phys. Rev. Lett. 98, 050405 (2007).
  • [7] L. Fleishman and P. W. Anderson, Interactions and the Anderson Transition, Phys. Rev. B 21, 2366 (1980).
  • [8] M. Žnidarič, T. Prosen, and P. Prelovšek, Many-Body Localization in the Heisenberg XXZXXZ Magnet in a Random Field, Phys. Rev. B 77, 064426 (2008).
  • [9] A. Pal and D. A. Huse, Many-Body Localization Phase Transition, Phys. Rev. B 82, 174411 (2010).
  • [10] M. Serbyn, Z. Papić, and D. A. Abanin, Universal Slow Growth of Entanglement in Interacting Strongly Disordered Systems, Phys. Rev. Lett. 110, 260601 (2013).
  • [11] D. A. Huse, R. Nandkishore, and V. Oganesyan, Phenomenology of Fully Many-Body-Localized Systems, Phys. Rev. B 90, 174202 (2014).
  • [12] J. Z. Imbrie, Diagonalization and Many-Body Localization for a Disordered Quantum Spin Chain, Phys. Rev. Lett. 117, 027201 (2016).
  • [13] J. Z. Imbrie, V. Ros, and A. Scardicchio, Local Integrals of Motion in Many-Body Localized Systems, ANNALEN DER PHYSIK 529, 1600278 (2017).
  • [14] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalattore, Colloquium: Nonequilibrium Dynamics of Closed Interacting Quantum Systems, Rev. Mod. Phys. 83, 863 (2011).
  • [15] 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).
  • [16] J. Choi, S. Hild, J. Zeiher, P. Schauß, A. Rubio-Abadal, T. Yefsah, V. Khemani, D. A. Huse, I. Bloch, and C. Gross, Exploring the Many-Body Localization Transition in Two Dimensions, Science 352, 1547 (2016).
  • [17] J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess, P. Hauke, M. Heyl, D. A. Huse, and C. Monroe, Many-Body Localization in a Quantum Simulator with Programmable Random Disorder, Nat. Phys. 12, 907 (2016).
  • [18] K. X. Wei, C. Ramanathan, and P. Cappellaro, Exploring Localization in Nuclear Spin Chains, Phys. Rev. Lett. 120, 070501 (2018).
  • [19] T. Orell, A. A. Michailidis, M. Serbyn, and M. Silveri, Probing the Many-Body Localization Phase Transition with Superconducting Circuits, Phys. Rev. B 100, 134504 (2019).
  • [20] C. Neill et al., A Blueprint for Demonstrating Quantum Supremacy with Superconducting Qubits, Science 360, 195 (2018).
  • [21] A. Smith, J. Knolle, D. L. Kovrizhin, and R. Moessner, Disorder-Free Localization, Phys. Rev. Lett. 118, 266601 (2017).
  • [22] M. Brenes, M. Dalmonte, M. Heyl, and A. Scardicchio, Many-Body Localization Dynamics from Gauge Invariance, Phys. Rev. Lett. 120, 030601 (2018).
  • [23] I. Papaefstathiou, A. Smith, and J. Knolle, Disorder-Free Localization in a Simple U(1)U(1) Lattice Gauge Theory, Phys. Rev. B 102, 165132 (2020).
  • [24] P. Karpov, R. Verdel, Y.-P. Huang, M. Schmitt, and M. Heyl, Disorder-Free Localization in an Interacting 2D Lattice Gauge Theory, Phys. Rev. Lett. 126, 130401 (2021).
  • [25] J. C. Halimeh, H. Zhao, P. Hauke, and J. Knolle, Stabilizing Disorder-Free Localization, arXiv: 2111.02427.
  • [26] H. Lang, P. Hauke, J. Knolle, F. Grusdt, and J. C. Halimeh, Disorder-Free Localization with Stark Gauge Protection, Phys. Rev. B 106, 174305 (2022).
  • [27] J. C. Halimeh, P. Hauke, J. Knolle, and F. Grusdt, Temperature-Induced Disorder-Free Localization, arXiv: 2206.11273.
  • [28] B. Yang, H. Sun, R. Ott, H.-Y. Wang, T. V. Zache, J. C. Halimeh, Z.-S. Yuan, P. Hauke, and J.-W. Pan, Observation of Gauge Invariance in a 71-Site Bose-Hubbard Quantum Simulator, Nature 587, 392 (2020).
  • [29] A. Mil, T. V. Zache, A. Hegde, A. Xia, R. P. Bhatt, M. K. Oberthaler, P. Hauke, J. Berges, and F. Jendrzejewski, A Scalable Realization of Local U(1) Gauge Invariance in Cold Atomic Mixtures, Science 367, 1128 (2020).
  • [30] Z.-Y. Zhou, G.-X. Su, J. C. Halimeh, R. Ott, H. Sun, P. Hauke, B. Yang, Z.-S. Yuan, J. Berges, and J.-W. Pan, Thermalization Dynamics of a Gauge Theory on a Quantum Simulator, Science 377, 311 (2022).
  • [31] C. Gao, J. Liu, M. Chang, H. Pu, and L. Chen, Synthetic U(1) Gauge Invariance in a Spin-1 Bose Gas, Phys. Rev. Research 4, L042018 (2022).
  • [32] J. Schwinger, The Theory of Quantized Fields. I, Phys. Rev. 82, 914 (1951).
  • [33] J. Schwinger, The Theory of Quantized Fields. II, Phys. Rev. 91, 713 (1953).
  • [34] S. Coleman, More about the Massive Schwinger Model, Annals of Physics 101, 239 (1976).
  • [35] J. Kogut and L. Susskind, Hamiltonian Formulation of Wilson Lattice Gauge Theories, Phys. Rev. D 11, 395 (1975).
  • [36] Y. Cheng, S. Liu, W. Zheng, P. Zhang, and H. Zhai, Tunable Confinement-Deconfinement Transition in an Ultracold Atom Quantum Simulator, PRX Quantum 3, 040317 (2022).
  • [37] J. C. Halimeh, I. P. McCulloch, B. Yang, and P. Hauke, Tuning the Topological θ\theta-Angle in Cold-Atom Quantum Simulators of Gauge Theories, PRX Quantum 3, 040316 (2022).
  • [38] S. Chandrasekharan, U.-J. Wiese, Quantum Link Models: A Discrete Approach to Gauge Theories, Nucl. Phys. B 492, 455-474 (1997).
  • [39] U.-J. Wiese, Ultracold Quantum Gases and Lattice Systems: Quantum Simulation of Lattice Gauge Theories, ANNALEN DER PHYSIK 525, 777 (2013).
  • [40] E. Zohar, J. I. Cirac, and B. Reznik, Simulating Compact Quantum Electrodynamics with Ultracold Atoms: Probing Confinement and Nonperturbative Effects, Phys. Rev. Lett. 109, 125302 (2012).
  • [41] E. Zohar, J. I. Cirac, and B. Reznik, Quantum Simulations of Lattice Gauge Theories Using Ultracold Atoms in Optical Lattices, Rep. Prog. Phys. 79, 014401 (2016).
  • [42] D. Banerjee, M. Dalmonte, M. Müller, E. Rico, P. Stebler, U.-J. Wiese, and P. Zoller, Atomic Quantum Simulation of Dynamical Gauge Fields Coupled to Fermionic Matter: From String Breaking to Evolution after a Quench, Phys. Rev. Lett. 109, 175302 (2012).
  • [43] G. Magnifico, M. Dalmonte, P. Facchi, S. Pascazio, F. V. Pepe, and E. Ercolessi, Real Time Dynamics and Confinement in the n\mathbb{Z}_{n} Schwinger-Weyl Lattice Model for 1+1 QED, Quantum 4, 281 (2020).
  • [44] F. M. Surace, P. P. Mazza, G. Giudici, A. Lerose, A. Gambassi, and M. Dalmonte, Lattice Gauge Theories and String Dynamics in Rydberg Atom Quantum Simulators, Phys. Rev. X 10, 021041 (2020).
  • [45] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres, M. Greiner, V. Vuletić, and M. D. Lukin, Probing Many-Body Dynamics on a 51-Atom Quantum Simulator, Nature 551, 579 (2017).
  • [46] C. J. Turner, A. A. Michailidis, D. A. Abanin, M. Serbyn, and Z. Papić, Weak Ergodicity Breaking from Quantum Many-Body Scars, Nat. Phys. 14, 745 (2018).
  • [47] M. Serbyn, D. A. Abanin, and Z. Papić, Quantum Many-Body Scars and Weak Breaking of Ergodicity, Nat. Phys. 17, 675 (2021).
  • [48] A. W. Sandvik, A. Avella, and F. Mancini, Computational Studies of Quantum Spin Systems, in (Vietri Sul Mare, (Italy), 2010).
  • [49] P. Weinberg and M. Bukov, QuSpin: A Python Package for Dynamics and Exact Diagonalisation of Quantum Many Body Systems Part I: Spin Chains, SciPost Physics 2, 003 (2017).
  • [50] N. Y. Yao, C. R. Laumann, J. I. Cirac, M. D. Lukin, and J. E. Moore, Quasi-Many-Body Localization in Translation-Invariant Systems, Phys. Rev. Lett. 117, 240601 (2016).
  • [51] M. Schiulaz, A. Silva, and M. Müller, Dynamics in Many-Body Localized Quantum Systems without Disorder, Phys. Rev. B 91, 184202 (2015).
  • [52] T. Enss, F. Andraschko, and J. Sirker, Many-Body Localization in Infinite Chains, Phys. Rev. B 95, 045121 (2017).
  • [53] M. Müller-Lennert, F. Dupuis, O. Szehr, S. Fehr, and M. Tomamichel, On Quantum Rényi Entropies: A New Generalization and Some Properties, Journal of Mathematical Physics 54, 122203 (2013).
  • [54] J. Berges, Sz. Borsányi, and C. Wetterich, Prethermalization, Phys. Rev. Lett. 93, 142002 (2004).