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

Lindbladian Many-Body Localization

Ryusuke Hamazaki Nonequilibrium Quantum Statistical Mechanics RIKEN Hakubi Research Team, RIKEN Cluster for Pioneering Research (CPR), RIKEN iTHEMS, Wako, Saitama 351-0198, Japan    Masaya Nakagawa Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan    Taiki Haga Department of Physics and Electronics, Osaka Metropolitan University, Osaka, 599-8531, Japan    Masahito Ueda Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan Institute for Physics of Intelligence, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
Abstract

We discover a novel localization transition that alters the dynamics of coherence in disordered many-body spin systems subject to Markovian dissipation. The transition occurs in the middle spectrum of the Lindbladian super-operator whose eigenstates obey the universality of non-Hermitian random-matrix theory for weak disorder and exhibit localization of off-diagonal degrees of freedom for strong disorder. This Lindbladian many-body localization prevents many-body decoherence due to interactions and is conducive to robustness of the coherent dynamics characterized by the rigidity of the decay rate of coherence.

Introduction.

Strongly disordered potentials of isolated systems significantly alter quantum dynamics. Spectral statistics of a Hamiltonian characterize thermalizing and non-thermalizing phases. For weak disorder, the eigenvalue-spacing distribution obeys the Wigner-Dyson statistics, and the eigenstate thermalization hypothesis (ETH) Srednicki (1994); Deutsch (1991); Rigol et al. (2008); Rigol (2009); Biroli et al. (2010); Genway et al. (2012); Khatami et al. (2013); Steinigeweg et al. (2013); Beugeling et al. (2014); Kim et al. (2014); Beugeling et al. (2015); Luitz and Bar Lev (2016); Yoshizawa et al. (2018); Hamazaki and Ueda (2019); Khaymovich et al. (2019); Jansen et al. (2019); Nation and Porras (2019); Sugimoto et al. (2021a); Fritzsch and Prosen (2021); Sugimoto et al. (2021b); D’Alessio et al. (2016); Mori et al. (2018) holds, reflecting the universality of Hermitian random matrix theory (RMT) Haake (2010). For strong disorder, eigenvalues obey the Poisson distribution, and eigenstates are described by quasi-local integrals of motion, providing unique features of many-body localization (MBL) Basko et al. (2006); Žnidarič et al. (2008); Pal and Huse (2010); Gogolin et al. (2011); Bardarson et al. (2012); Iyer et al. (2013); Huse et al. (2013); Serbyn et al. (2013); Huse et al. (2014); Kjäll et al. (2014); Nandkishore and Huse (2015); Bera et al. (2015); Ponte et al. (2015); Potter et al. (2015); Schreiber et al. (2015); Luitz et al. (2015); Vosk et al. (2015); Smith et al. (2016); Choi et al. (2016); Imbrie (2016a, b); Khemani et al. (2017); Alet and Laflorencie (2018); Lukin et al. (2019); Abanin et al. (2019).

However, no system is immune to external dissipation Syassen et al. (2008); Barreiro et al. (2011); Yan et al. (2013); Barontini et al. (2013); Labouvie et al. (2015); Patil et al. (2015); Gao et al. (2015); Labouvie et al. (2016); Lüschen et al. (2017); Raftery et al. (2014); Tomita et al. (2017); Lapp et al. (2019); Takasu et al. (2020); Bouganne et al. (2020); Keßler et al. (2021); Noel et al. (2021), which dramatically alters the nature of MBL Nandkishore et al. (2014); Johri et al. (2014); Fischer et al. (2016); Levi et al. (2016); Medvedyeva et al. (2016); Lüschen et al. (2017); van Nieuwenburg et al. (2017); Morningstar et al. (2022); Sels (2021). By coupling a small bath Hamiltonian to an MBL Hamiltonian, the authors of Refs. Nandkishore et al. (2014); Johri et al. (2014) discuss that the signature of MBL on the spectral functions of spins survives under dissipation; however, the total system becomes delocalized and satisfies the ETH. Instead, the authors of Refs. Fischer et al. (2016); Levi et al. (2016); Medvedyeva et al. (2016); Lüschen et al. (2017) incorporate dissipation through the Lindblad equation and show that the local integrals of motion are no longer preserved, rendering the stationary state delocalized. These works mainly focus on how the signature of the Hermitian MBL is altered by dissipation. However, a question remains concerning whether a sharp localization transition exists as a unique phenomenon for a dissipative disordered many-body system. For example, is there a spectral transition described by the Lindbladian super-operator as in the Hamiltonian operator in the Hermitian MBL pre , and, if any, what is the consequence of the transition on the dynamics? Note that the non-Hermitian MBL Hamazaki et al. (2019) describes the post-selected dynamics of continuously measured systems via an effective non-Hermitian Hamiltonian and is inapplicable to generic Lindbladian dynamics.

Refer to caption
Figure 1: Phase diagram of a dissipative many-body system with disorder. For finite dissipation γ>0\gamma>0 and sufficiently strong disorder hh, the oscillation of an observable O^q\hat{O}_{q} corresponding to quantum coherence attenuates slowly (red solid curve in the right inset), whereas for weak disorder it decays rapidly (blue solid curve in the left inset). Note that the decay rate is stabilized due to the Lindbladian many-body localization (see the main text). These two regimes are distinguished by the spectral statistics featuring the localization and non-Hermitian RMT of the Lindbladian super-operator, as a dissipative generalization of the Hermitian Hamiltonian case (γ=0\gamma=0). The non-Hermitian RMT classes depend on the type of symmetry of the Lindblad dynamics that is determined by dephasing or damping.

In this work, we show that an unconventional localization occurs for dissipative systems as a spectral transition of the Lindbladian super-operator (rather than the Hamiltonian), which alters the coherent dynamics. In fact, spectral statistics in the middle of the spectrum are characterized by the universality of the non-Hermitian RMT for weak disorder and by the localization of the Lindbladian eigenstates for strong disorder (see Fig. 1). We call the latter unique open dissipative localization as the Lindbladian MBL. Deep in the Lindbladian MBL regime, quantum coherence exhibits a rigid decay whose rate is essentially determined by the decoherence rate for a single-spin system and stabilized at a roughly constant value against the variations of interaction terms and the magnetic field. This rigidity of the decay rate is attributed to localization of off-diagonal degrees of freedom in the Lindbladian eigenstates. The behavior is distinct from interaction-induced many-body decoherence for weak disorder, where eigenmodes with many different frequencies and decay rates govern the dynamics. Using a prototypical Ising model with a dephasing- or damping-type of dissipation, we numerically demonstrate the transition with, e.g., the complex spacing distributions of eigenvalues and the operator-space entanglement entropy (OSEE) of eigenstates.

Dissipative spin chains with disorder.

We consider a one-dimensional Ising model with transverse and longitudinal fields under dissipation. The dynamics is described by the Lindblad equation with =1\hbar=1 Lindblad (1976):

dρ^dt=[ρ^]=i[H^,ρ^]+l12[2Γ^lρ^Γ^l{Γ^lΓ^l,ρ^}].\displaystyle\frac{\mathrm{d}\hat{\rho}}{\mathrm{d}t}=\mathcal{L}[\hat{\rho}]=-i[\hat{H},\hat{\rho}]+\sum_{l}\frac{1}{2}\left[2\hat{\Gamma}_{l}\hat{\rho}\hat{\Gamma}_{l}^{\dagger}-{\left\{\hat{\Gamma}_{l}^{\dagger}\hat{\Gamma}_{l},\hat{\rho}\right\}}\right]. (1)

Here, the Hamiltonian is given by H^=i=1L1Jσ^izσ^i+1z+i=1Lgσ^ix+i=1Lhiσ^iz,\hat{H}=\sum_{i=1}^{L-1}J\hat{\sigma}_{i}^{z}\hat{\sigma}_{i+1}^{z}+\sum_{i=1}^{L}g\hat{\sigma}_{i}^{x}+\sum_{i=1}^{L}h_{i}\hat{\sigma}_{i}^{z}, where hih_{i} is taken randomly from [h,+h][-h,+h]. Without dissipation, H^\hat{H} exhibits MBL for sufficiently strong disorder Imbrie (2016a, b). We consider two types of dissipation, namely dephasing Γ^l=γσ^lz,\hat{\Gamma}_{l}=\sqrt{\gamma}\hat{\sigma}^{z}_{l}, and damping Γ^l=γ2σ^l\hat{\Gamma}_{l}={\sqrt{\frac{\gamma}{2}}}\hat{\sigma}^{-}_{l} with σ^l=σ^lxiσ^ly\hat{\sigma}^{-}_{l}=\hat{\sigma}^{x}_{l}-i\hat{\sigma}^{y}_{l}. Below we mainly consider the case with weak dissipation (γ<g\gamma<g). The case with large dissipation (γ>g)(\gamma>g) is discussed at the Discussion section.

We first discuss the time evolution of coherence measured by O^q=k=Lq+1Lσ^kx\hat{O}_{q}=\bigotimes_{k=L-q+1}^{L}\hat{\sigma}_{k}^{x} starting from an initial state ρ^q=|ψqψq|\hat{\rho}_{q}=\ket{\psi_{q}}\bra{\psi_{q}}, where |ψq=(|1|Lq)(|Lq+1|L+|Lq+1|L2).\ket{\psi_{q}}=(\ket{\downarrow}_{1}\cdots\ket{\downarrow}_{L-q})\left(\frac{\ket{\uparrow}_{L-q+1}\cdots\ket{\uparrow}_{L}+\ket{\downarrow}_{L-q+1}\cdots\ket{\downarrow}_{L}}{\sqrt{2}}\right). This choice corresponds to measuring the single-spin coherence for q=1q=1 and macroscopic coherence of the Greenberger-Horne-Zeilinger state for q=Lq=L. In Fig. 2(a), we show the time evolution of the quantum expectation value of O^q(t)\braket{\hat{O}_{q}(t)} (not averaged over disorder) for different values of disorder strength. While O^q(t)\braket{\hat{O}_{q}(t)} rapidly relaxes to a stationary value for small hh, it exhibits a slower oscillatory decay for large hh. Only for the latter case, the decay rate is almost stabilized at around 2γq(γq)2\gamma q~{}(\gamma q) for dephasing (damping) pol , which is explained by the decoherence rate for a single-spin system and robust against many-body interactions (see Supplemental Materials Sup ).

Refer to caption
Figure 2: (a) Time evolution of O^q(t)\braket{\hat{O}_{q}(t)} for h=1.2h=1.2 (blue) and h=10h=10 (red). While O^q(t)\braket{\hat{O}_{q}(t)} rapidly relaxes to a stationary state for small hh, it exhibits a slower oscillatory decay for large hh. The orange dashed curves show ±e2qγt\pm e^{-2q\gamma t} for the dephasing and ±eqγt\pm e^{-q\gamma t} for the damping type of dissipation, which approximate the amplitudes of the decay for large hh. (b) The distribution of |Fα||F_{\alpha}| as a function of the eigenvalue λα\lambda_{\alpha} for the dephasing-type dissipation. While |Fα||F_{\alpha}| in Eq. (2) spreads over many α\alpha’s for h=1.2h=1.2 (blue dots), peaks appear for several eigenstates embedded in the middle of the spectrum, and the other |Fα||F_{\alpha}| are vanishingly small for h=10h=10 (red asterisks). We use L=6,J=1,g=0.9,γ=0.1L=6,J=1,g=-0.9,\gamma=0.1 and q=3q=3.

To understand the distinction between two regimes, we consider the spectral decomposition of the dynamics Gong and Hamazaki (2022):

O^q(t)\displaystyle\braket{\hat{O}_{q}(t)} =αTr[O^qR^α]Tr[L^αρ^q]Tr[L^αR^α]eλαt=:αFαeλαt,\displaystyle=\sum_{\alpha}\frac{\mathrm{Tr}[\hat{O}_{q}\hat{R}_{\alpha}]\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}_{q}]}{\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{R}_{\alpha}]}e^{\lambda_{\alpha}t}=:\sum_{\alpha}F_{\alpha}e^{\lambda_{\alpha}t}, (2)

where R^α\hat{R}_{\alpha} (L^α\hat{L}_{\alpha}) is a right (left) eigenstate of the Lindbladian super-operator \mathcal{L} with an eigenvalue λα\lambda_{\alpha}\in\mathbb{C} nor . For weak disorder, many eigenstates for different λα\lambda_{\alpha} contribute to the dynamics, i.e., |Fα||F_{\alpha}| in Eq. (2) spreads over many α\alpha (see Fig. 2(b)), indicating complicated many-body decoherence characterized by a large number of frequencies and decay rates. On the other hand, for strong disorder, peaks in |Fα||F_{\alpha}| appear for several eigenstates in the middle of the spectrum with the other |Fα||F_{\alpha}| being vanishingly small. These selected eigenstates lead to an oscillatory decay governed by only a few frequencies and a rigid decay rate despite many-body interactions.

Non-Hermitian random-matrix universality and many-body decoherence.

The distinctive behavior of the behavior of |Fα||F_{\alpha}| discussed above is attributed to the different regimes characterized by the spectral statistics of the Lindbladian super-operator. For weak disorder, we find that the statistics obey the non-Hermitian RMT. In particular, the trace factors appearing in Eq. (2) in the middle of the spectrum and away from the real axis are described as

Tr[O^R^α]𝒜O^(λα)rα,\displaystyle\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]\sim\mathcal{A}_{\hat{O}}(\lambda_{\alpha})r_{\alpha}, (3)
Tr[L^αρ^]ρ^(λα)rα,\displaystyle\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}]\sim\mathcal{B}_{\hat{\rho}}(\lambda_{\alpha})r_{\alpha}, (4)

and

Tr[L^αR^α]1𝒞(λα)rα,\displaystyle\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{R}_{\alpha}]^{-1}\sim\mathcal{C}(\lambda_{\alpha})r^{\prime}_{\alpha}, (5)

where 𝒜O^\mathcal{A}_{\hat{O}}, ρ^\mathcal{B}_{\hat{\rho}}, and 𝒞\mathcal{C} are smooth functions of λα\lambda_{\alpha}, and rαr_{\alpha} and rαr^{\prime}_{\alpha} are complex random variables normalized as 𝔼[|rα|]=𝔼[|rα|]=1\mathbb{E}[|r_{\alpha}|]=\mathbb{E}[|r_{\alpha}^{\prime}|]=1. These forms indicate that every energy eigenstate fluctuates randomly within a sufficiently small two-dimensional eigenvalue window where 𝒜O^(λα)\mathcal{A}_{\hat{O}}(\lambda_{\alpha}), ρ^(λα)\mathcal{B}_{\hat{\rho}}(\lambda_{\alpha}), and 𝒞(λα)\mathcal{C}(\lambda_{\alpha}) are almost constant. In this sense, Eqs. (3)-(5) constitute a dissipative generalization of the (off-diagonal) ETH for Hermitian systems, which states that Ea|O^|Eb\braket{E_{a}}{\hat{O}}{E_{b}} fluctuates within the small energy windows around EaE_{a} and EbE_{b} according to (Hermitian) RMT Khatami et al. (2013) for quantum chaotic systems cla .

Notably, the distributions of rαr_{\alpha} and rαr^{\prime}_{\alpha} obey the non-Hermitian RMT universality Ginibre (1965); Chalker and Mehlig (1998); Mehlig and Chalker (2000). Specifically, the distributions P(rα)P(r_{\alpha}) and P(rα)P(r^{\prime}_{\alpha}) are described by PG(x):=P(|rαG|=x)=π2xeπ2x2,P_{G}(x):=P(|r_{\alpha}^{G}|=x)=\frac{\pi}{2}xe^{-\frac{\pi}{2}x^{2}}, which is obtained from the Porter-Thomas distribution, and PG(x):=P(|rαG|=x)=32π2x5e4πx2,P_{G}^{\prime}(x):=P(|r_{\alpha}^{G^{\prime}}|=x)=\frac{32}{\pi^{2}x^{5}}e^{-\frac{4}{\pi x^{2}}}, which is obtained from the results of a complex Ginibre ensemble Bourgade and Dubach (2019); rea . As shown in Fig. 3, these results are valid for our model with weak disorder, but invalid for strong disorder pea .

Refer to caption
Figure 3: Distributions of the normalized values of |Tr[L^αR^α]|1,|Tr[L^αρ^]|,|\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{R}_{\alpha}]|^{-1},|\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}]|, and |Tr[O^R^α]||\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]|, i.e., rαr_{\alpha}^{\prime} and rαr_{\alpha}, for h=1.2h=1.2 (top) and h=4.0h=4.0 (bottom). We see that P(|rα|=x)P(|r^{\prime}_{\alpha}|=x) and P(|rα|=x)P(|r_{\alpha}|=x) obey non-Hermitian random-matrix distributions PGP^{\prime}_{G} and PGP_{G}, respectively, for sufficiently small hh. The distributions deviate from the non-Hermitian random-matrix predictions for large hh. We use q=1q=1, L=7,J=1,g=0.9L=7,J=1,g=-0.9, and γ=0.5\gamma=0.5 with dephasing-type dissipation. The distributions are calculated from eigenstates within a small range of eigenvalues with 50 ensembles.

As detailed in Supplemental Material Sup , |Tr[O^R^α]||\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]| and |Tr[L^αρ^]||\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}]| decrease, and |Tr[L^αR^α]|1|\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{R}_{\alpha}]|^{-1} increases with increasing LL. Furthermore, |Tr[L^αR^α]|1|\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{R}_{\alpha}]|^{-1} for small hh is almost proportional to the dimension D=2LD=2^{L} of the Hilbert space. We also find Tr[L^αρ^]D1\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}]\propto D^{-1} for small hh. These scalings agree with the prediction of the non-Hermitian RMT. On the other hand, for strong hh, the scaling behavior differs from the RMT for both cases. Note that the scaling behavior is not simple for |Tr[O^R^α]||\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]| because of the locality of the operator and that of the Lindbladian. In this case, we argue that |Tr[O^R^α]|ec|λα|(c>0)|\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]|\sim e^{-c^{\prime}|\lambda_{\alpha}|}\>(c^{\prime}>0) Sup .

The above discussions indicate that, for weak disorder, eigenstates within a small eigenvalue window fluctuate randomly for |Tr[R^αL^α]|1,|Tr[L^αρ^]|,|\mathrm{Tr}[\hat{R}_{\alpha}^{\dagger}\hat{L}_{\alpha}]|^{-1},|\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}]|, and |Tr[O^R^α]||\mathrm{Tr}[\hat{O}\hat{R}_{\alpha}]|. Consequently, FαF_{\alpha} also behaves randomly without irregular eigenstates as shown in Fig. 2(b), resulting in many-body decoherence governed by many modes with different frequencies and decay rates. The above discussion establishes the previously unknown connection between dissipative quantum chaos and many-body decoherence in terms of eigenstates with the RMT universality.

Lindbladian MBL.

We next discuss the strong-disorder case. We start from the phenomenological picture with quasi-local bits Huse et al. (2014) in the MBL phase without dissipation. Then, the Hamiltonian reads

H^=lh~lτ^lz+lmJlmτ^lzτ^mz+lmnJlmnτ^lzτ^mzτ^nz+,\displaystyle\hat{H}=\sum_{l}\tilde{h}_{l}\hat{\tau}_{l}^{z}+\sum_{lm}J_{lm}\hat{\tau}_{l}^{z}\hat{\tau}_{m}^{z}+\sum_{lmn}J_{lmn}\hat{\tau}_{l}^{z}\hat{\tau}_{m}^{z}\hat{\tau}_{n}^{z}+\cdots, (6)

where h~lhl\tilde{h}_{l}\simeq h_{l}, and Jlm,Jlmn,J_{lm},J_{lmn},\cdots decay exponentially for any far apart two-site indices. The local integral of motion τ^lz\hat{\tau}_{l}^{z} has a large overlap with σ^lz\hat{\sigma}_{l}^{z}. The Hamiltonian H^\hat{H} has eigenstates labeled by the eigenvalues ±1\pm 1 of τ^lz\hat{\tau}_{l}^{z}, i.e., |τ1τL\ket{\tau_{1}\cdots\tau_{L}}. We can express σ^l\hat{\sigma}_{l} by τ^l\hat{\tau}_{l} as σ^lz=α=x,y,zZlατ^lα+jkα,β=x,y,zGl,jkαβτ^jατ^kβ+(higher order terms)\hat{\sigma}_{l}^{z}=\sum_{\alpha=x,y,z}Z_{l}^{\alpha}\hat{\tau}_{l}^{\alpha}+\sum_{jk}\sum_{\alpha,\beta=x,y,z}G_{l,jk}^{\alpha\beta}\hat{\tau}_{j}^{\alpha}\hat{\tau}_{k}^{\beta}+\text{(higher order terms)}, where Zlz=1O((g/h)2)Z_{l}^{z}=1-\mathrm{O}((g/h)^{2}), Zlx,y=O(g/h)Z_{l}^{x,y}=\mathrm{O}(g/h), and Gl,jkαβG_{l,jk}^{\alpha\beta} rapidly decays as a function of |lj||l-j| and |lk||l-k|. Also, Gl,jkαβ=O((g/h)2)G_{l,jk}^{\alpha\beta}=\mathrm{O}((g/h)^{2}) Imbrie (2016a, b). A similar representation is obtained for σ^l\hat{\sigma}_{l}^{-}. Then, the Lindbladian reads [ρ^]=i[H^,ρ^]+Z+P\mathcal{L}[\hat{\rho}]=-i[\hat{H},\hat{\rho}]+\mathcal{L}_{Z}+\mathcal{L}_{P}, where Z=Zlz22l[2Γ^lρ^Γ^l{Γ^lΓ^l,ρ^}]\mathcal{L}_{Z}=\frac{{Z_{l}^{z}}^{2}}{2}\sum_{l}\left[2\hat{\Gamma}_{l}^{\prime}\hat{\rho}\hat{\Gamma}_{l}^{\prime{\dagger}}-{\left\{\hat{\Gamma}_{l}^{\prime{\dagger}}\hat{\Gamma}_{l}^{\prime},\hat{\rho}\right\}}\right] with Γ^l=γτ^lz\hat{\Gamma}_{l}^{\prime}=\sqrt{\gamma}\hat{\tau}_{l}^{z} (dephasing) or Γ^l=γ2τ^z\hat{\Gamma}_{l}^{\prime}=\sqrt{\frac{\gamma}{2}}\hat{\tau}_{-}^{z} (damping), and P\mathcal{L}_{P} denotes the remaining perturbation.

We briefly discuss the case of dephasing-type dissipation here (see Supplemental Material for details and the case of damping-type dissipation Sup ). The eigenstates of i[H^,ρ^]+Z-i[\hat{H},\hat{\rho}]+\mathcal{L}_{Z} can be written as R^α=L^α=|τ1τLτ1τL|=ϕ^1α1ϕ^2α2ϕ^LαL\hat{R}_{\alpha}=\hat{L}_{\alpha}=\ket{\tau_{1}\cdots\tau_{L}}\bra{\tau_{1}^{\prime}\cdots\tau_{L}^{\prime}}=\hat{\phi}_{1}^{\alpha_{1}}\otimes\hat{\phi}_{2}^{\alpha_{2}}\otimes\cdots\otimes\hat{\phi}_{L}^{\alpha_{L}}. Here, \otimes represents the tensor product for different localized bits τ^i\hat{\tau}_{i}, and ϕ^lαl=|τlτl|\hat{\phi}_{l}^{\alpha_{l}}=\ket{\tau_{l}}\bra{\tau_{l}^{\prime}}. Specifically, we take ϕ^1=(ϕ^2)=|+11|\hat{\phi}^{1}=(\hat{\phi}^{2})^{\dagger}=\ket{+1}\bra{-1}, ϕ^3=|+1+1|\hat{\phi}^{3}=\ket{+1}\bra{+1}, and ϕ^4=|11|\hat{\phi}^{4}=\ket{-1}\bra{-1}, where |±1\ket{\pm 1} is the eigenstate of τ^z\hat{\tau}^{z} with an eigenvalue ±1\pm 1. The corresponding eigenvalues are il2h~l(δαl,1δαl,2)i(small terms including Jlm,)l2Zlz2γ(δαl,1+δαl,2)-i\sum_{l}2\tilde{h}_{l}(\delta_{\alpha_{l},1}-\delta_{\alpha_{l},2})-i(\text{small terms including }J_{lm},\cdots)-\sum_{l}2{Z_{l}^{z}}^{2}\gamma(\delta_{\alpha_{l},1}+\delta_{\alpha_{l},2}).

When we add P\mathcal{L}_{P}, ϕ^l3\hat{\phi}_{l}^{3} and ϕ^l4\hat{\phi}_{l}^{4} in different eigenstates are mixed, since the eigenvalue difference is almost zero in this case. In contrast, ϕ^l1\hat{\phi}_{l}^{1} and ϕ^l2\hat{\phi}_{l}^{2} are typically stable under first-order perturbation, since the transition matrix elements over the eigenvalue difference are O(γg/h2)1\mathrm{O}(\gamma g/h^{2})\ll 1 Sup . It follows then that while diagonal degrees of freedom (DDOF) ϕ^l3\hat{\phi}_{l}^{3} and ϕ^l4\hat{\phi}_{l}^{4} can delocalize, off-diagonal degrees of freedom (ODDOF) ϕ^l1\hat{\phi}_{l}^{1} and ϕ^l2\hat{\phi}_{l}^{2} can localize. Consequently, eigenstates are not fully mixed, and the RMT prediction breaks down. In the dynamics of coherence considered in Fig. 2 Coh , only the modes with αLq+1==αL=1\alpha_{L-q+1}=\cdots=\alpha_{L}=1 or 2 contribute to FαF_{\alpha}, and the other modes make negligible contributions to FαF_{\alpha}. The decay rate is then stabilized, and the system evades many-body decoherence. In particular, the decay rate becomes 2qγ(1O((g/h)2))\sim 2q\gamma(1-\mathrm{O}((g/h)^{2})). For the damping case, the decay rate is similarly stabilized at qγ(1O((g/h)2))\sim q\gamma(1-\mathrm{O}((g/h)^{2})).

The stabilized dynamics of the coherence (transverse relaxation) is understood from the nontrivial Lindbladian localization of the ODDOF, which cannot be captured by the classical effective rate equation Fischer et al. (2016); Medvedyeva et al. (2016) used for describing the slow longitudinal relaxation in previous studies. Indeed, the DDOF that govern the longitudinal relaxation Fischer et al. (2016); Medvedyeva et al. (2016) are delocalized. A detailed discussion of the difference between longitudinal and transverse relaxation is given in Supplemental Material Sup .

A few remarks are in order here. First, the delocalization/localization of DDOF/ODDOF is reasonable because they undergo zero/strong random fields 0/±2h~l\sim 0/\pm 2\tilde{h}_{l} in the matrix representation of the Lindbladian Sup . Second, the localization of ODDOF indicates the emergence of a quasi-local weak symmetry Buča and Prosen (2012); Albert and Jiang (2014) U^w\hat{U}_{w} of the Lindbladian, i.e., U^w[ρ^]U^w=[U^wρ^U^w]\hat{U}_{w}\mathcal{L}[\hat{\rho}]\hat{U}_{w}^{\dagger}=\mathcal{L}[\hat{U}_{w}\hat{\rho}\hat{U}_{w}^{\dagger}], which block-diagonalizes the Lindbladian. Third, as for the discussion of the Hermitian MBL, it is not easy to show the existence of localization and the precise transition point when we consider higher-order perturbations and resonant states Šuntajs et al. (2020); Morningstar et al. (2022); Sels (2021). We leave it as a future problem to investigate larger system sizes.

Delocalization-localization transition.

Refer to caption
Figure 4: (a) Complex spacing ratio rr as a function of disorder strength hh with L=6L=6 (green) and L=7L=7 (blue) for dephasing and damping types of dissipation. For small hh, rr becomes close to rAI0.721r_{\mathrm{AI^{\dagger}}}\simeq 0.721 (dashed black line) and rA0.737r_{\mathrm{A}}\simeq 0.737 (dashed orange line) for dephasing-type and damping-type, respectively. The value of rr decreases with increasing hh, and becomes close to that for the Poisson distribution rPo=2/3r_{\mathrm{Po}}=2/3 (dashed gray line). (b) System-size dependence of OSEE SαS_{\alpha} averaged over eigenstates in the middle of the spectrum for different disorder strength h=0.6,1.2,1.8,2.4,4.0,5.0,10.0h=0.6,1.2,1.8,2.4,4.0,5.0,10.0 (from red to orange). While the OSEE increases with increasing LL for small hh, the increase is suppressed for large hh. We use the dephasing type of dissipation. (c) Dependence on hh of the variance σ\sigma of the OSEE for 4L74\leq L\leq 7. The peak develops as LL increases, from which the transition point reads as hc=2.5±0.1(2.7±0.1)h_{c}=2.5\pm 0.1\>(2.7\pm 0.1) for dephasing (damping). For (a)-(c), we use J=1,g=0.9,J=1,g=-0.9, and γ=0.5\gamma=0.5 and the eigenvalues in the middle of the spectrum Mid . The error bars indicate standard errors of the quantities evaluated from samples with different disorder realizations. The number of samples is 10000(L=4,5),800(L=6),10000~{}(L=4,5),800~{}(L=6), and 50(L=7)50~{}(L=7).

To strengthen evidence for the Lindbladian MBL transition, we calculate the spacing statistics of complex eigenvalues Grobe et al. (1988); Grobe and Haake (1989); Markum et al. (1999); Akemann et al. (2019); Hamazaki et al. (2019, 2020); Jaiswal et al. (2019); Wang et al. (2020); Sá et al. (2020a, b); Tzortzakakis et al. (2020); Huang and Shklovskii (2020); Mudute-Ndumbe and Graefe (2020); Luo et al. (2021); Sá et al. (2021); Rubio-García et al. (2022); Prasad et al. (2022); García-García et al. (2022), particularly the complex spacing ratio r=𝔼α[|λαλαNλαλαNN|]r=\mathbb{E}_{\alpha}\left[\left|\frac{\lambda_{\alpha}-\lambda_{\alpha}^{\mathrm{N}}}{\lambda_{\alpha}-\lambda_{\alpha}^{\mathrm{NN}}}\right|\right] Sá et al. (2020a) as a function of hh in Fig. 4(a), where λαN\lambda_{\alpha}^{\mathrm{N}} and λαNN\lambda_{\alpha}^{\mathrm{NN}} are the nearest and the next-nearest neighbor eigenvalues of λα\lambda_{\alpha} on the complex plane. For small hh, rr is close to rAI0.721r_{\mathrm{AI^{\dagger}}}\simeq 0.721 and rA0.737r_{\mathrm{A}}\simeq 0.737 for dephasing-type and damping-type, respectively. Here, rAIr_{\mathrm{AI^{\dagger}}} and rAr_{\mathrm{A}} denote the complex spacing ratio for the non-Hermitian random matrices belonging to classes AI\mathrm{AI}^{\dagger} and A\mathrm{A}. On the other hand, rr decreases with increasing hh and eventually approaches the value for the Poisson distribution, rPo=2/3r_{\mathrm{Po}}=2/3 str . Note that Ref. Yusipov and Ivanchenko (2021) reports a similar eigenvalue transition for different disordered models, but with no mention to eigenstates and the change in the dynamics associated with localization.

We next consider the operator-space entanglement entropy (OSEE) SαS_{\alpha} Prosen and Pižorn (2007); Pižorn and Prosen (2009) of the (right) eigenstate R^α\hat{R}_{\alpha} for bipartition at the middle of the system l=L/2l=\lfloor L/2\rfloor OSE . Figure 4(b) shows the system-size dependence of SαS_{\alpha} averaged over eigenstates in the middle of the spectrum for different disorder strengths. While the OSEE increases with increasing LL for small hh, its shows a much slower increase for large hh. This is similar to the entanglement transition of the MBL in a closed system.

To find the transition point, we define the variance σ\sigma of the OSEE with respect to the eigenstates and show its disorder-strength dependence for various LL (Fig. 3(c)). The peak develops as we increase the system size, from which the transition point reads as hc=2.5±0.1(2.7±0.1)h_{c}=2.5\pm 0.1\>(2.7\pm 0.1) for dephasing (damping). The peak of σ\sigma at the transition is a dissipative counterpart of the peak of the variance of the entanglement entropy for the isolated MBL Kjäll et al. (2014).

Discussion.

We have discussed the transition for fixed γ\gamma with 0<γ<g0<\gamma<g. In contrast, as detailed in Supplemental Material Sup , by changing γ\gamma (including γg\gamma\geq g), we numerically find that the critical value hc(γ)h_{c}(\gamma) exhibits non-monotonic behavior. Namely, hc(γ)h_{c}(\gamma) first increases and then decreases. This implies that strong dissipation facilitates the Lindbladian MBL. For sufficiently large γ\gamma, we find the breakdown of the non-Hermitian RMT statistics even for small hh, as shown in Fig. 1. This indicates that a localized regime appears even for the clean system; however, we leave it as a future problem to investigate this possibility for larger system sizes.

In the opposite limit γ0\gamma\rightarrow 0, it is nontrivial whether the Lindbladian MBL transition point hc(γ0)h_{c}(\gamma\rightarrow 0) coincides with the Hermitian MBL transition point hH,ch_{\mathrm{H},c}. We conjecture that these two transition points coincide under certain conditions (see Supplemental Material Sup ).

Conclusion and outlook.

We have demonstrated that localization of Lindbladian eigenstates can occur in the open quantum many-body systems and stabilizes the decay rate of coherence without many-body decoherence despite interactions, if disorder is sufficiently strong. The weakly disordered phase is characterized by the spectral statistics reflecting the universality of non-Hermitian RMT, and the strongly disordered phase is characterized by the Lindbladian MBL.

Our study raises interesting questions. The nature of the transitions between the RMT and the Lindbladian MBL phases using larger system sizes needs to be investigated, which may uncover, e.g., a new type of criticality defined by the spectral statistics of super-operators in dissipative systems. It is also interesting to ask whether the delocalization-localization transitions of the spectrum found in this paper occur for models with different types of Hamiltonians (e.g., particle-number-conserving models) and dissipation (e.g., stochastic hopping Diehl et al. (2008); Yusipov et al. (2017); Vakulchyk et al. (2018); Haga et al. (2021)). Furthermore, it is intriguing to clarify the relation between our phases defined by the middle of the spectrum (relevant for transient-time dynamics) and the other types of MBL phenomenology under dissipation, such as the long-time thermalization Fischer et al. (2016); Levi et al. (2016); Medvedyeva et al. (2016); Lüschen et al. (2017); van Nieuwenburg et al. (2017); Morningstar et al. (2022); Sels (2021) and the non-Hermitian MBL Hamazaki et al. (2019).

Acknowledgements.
The numerical calculations were carried out with the help of QUSPIN Weinberg and Bukov (2017, 2019). M.N. is supported by JSPS KAKENHI Grant No. JP20K14383. T.H. is supported by JSPS KAKENHI Grant No. JP19J00525. M.U. is supported by JSPS KAKENHI Grant No. JP22H01152.

References

  • Srednicki (1994) M. Srednicki, Phys. Rev. E 50, 888 (1994).
  • Deutsch (1991) J. M. Deutsch, Phys. Rev. A 43, 2046 (1991).
  • Rigol et al. (2008) M. Rigol, V. Dunjko,  and M. Olshanii, Nature 452, 854 (2008).
  • Rigol (2009) M. Rigol, Phys. Rev. A 80, 053607 (2009).
  • Biroli et al. (2010) G. Biroli, C. Kollath,  and A. M. Läuchli, Phys. Rev. Lett. 105, 250401 (2010).
  • Genway et al. (2012) S. Genway, A. F. Ho,  and D. K. K. Lee, Phys. Rev. A 86, 023609 (2012).
  • Khatami et al. (2013) E. Khatami, G. Pupillo, M. Srednicki,  and M. Rigol, Phys. Rev. Lett. 111, 050403 (2013).
  • Steinigeweg et al. (2013) R. Steinigeweg, J. Herbrych,  and P. Prelovšek, Phys. Rev. E 87, 012118 (2013).
  • Beugeling et al. (2014) W. Beugeling, R. Moessner,  and M. Haque, Phys. Rev. E 89, 042112 (2014).
  • Kim et al. (2014) H. Kim, T. N. Ikeda,  and D. A. Huse, Phys. Rev. E 90, 052105 (2014).
  • Beugeling et al. (2015) W. Beugeling, R. Moessner,  and M. Haque, Phys. Rev. E 91, 012144 (2015).
  • Luitz and Bar Lev (2016) D. J. Luitz and Y. Bar Lev, Phys. Rev. Lett. 117, 170404 (2016).
  • Yoshizawa et al. (2018) T. Yoshizawa, E. Iyoda,  and T. Sagawa, Phys. Rev. Lett. 120, 200604 (2018).
  • Hamazaki and Ueda (2019) R. Hamazaki and M. Ueda, Phys. Rev. E 99, 042116 (2019).
  • Khaymovich et al. (2019) I. M. Khaymovich, M. Haque,  and P. A. McClarty, Phys. Rev. Lett. 122, 070601 (2019).
  • Jansen et al. (2019) D. Jansen, J. Stolpp, L. Vidmar,  and F. Heidrich-Meisner, Phys. Rev. B 99, 155130 (2019).
  • Nation and Porras (2019) C. Nation and D. Porras, Phys. Rev. E 99, 052139 (2019).
  • Sugimoto et al. (2021a) S. Sugimoto, R. Hamazaki,  and M. Ueda, Phys. Rev. Lett. 126, 120602 (2021a).
  • Fritzsch and Prosen (2021) F. Fritzsch and T. c. v. Prosen, Phys. Rev. E 103, 062133 (2021).
  • Sugimoto et al. (2021b) S. Sugimoto, R. Hamazaki,  and M. Ueda, arXiv preprint arXiv:2111.12484  (2021b).
  • D’Alessio et al. (2016) L. D’Alessio, Y. Kafri, A. Polkovnikov,  and M. Rigol, Advances in Physics 65, 239 (2016).
  • Mori et al. (2018) T. Mori, T. N. Ikeda, E. Kaminishi,  and M. Ueda, Journal of Physics B: Atomic, Molecular and Optical Physics 51, 112001 (2018).
  • Haake (2010) F. Haake, Quantum signatures of chaos, Vol. 54 (Springer Science & Business Media, 2010).
  • Basko et al. (2006) D. Basko, I. Aleiner,  and B. Altshuler, Annals of physics 321, 1126 (2006).
  • Žnidarič et al. (2008) M. Žnidarič, T. Prosen,  and P. Prelovšek, Phys. Rev. B 77, 064426 (2008).
  • Pal and Huse (2010) A. Pal and D. A. Huse, Phys. Rev. B 82, 174411 (2010).
  • Gogolin et al. (2011) C. Gogolin, M. P. Müller,  and J. Eisert, Phys. Rev. Lett. 106, 040401 (2011).
  • Bardarson et al. (2012) J. H. Bardarson, F. Pollmann,  and J. E. Moore, Phys. Rev. Lett. 109, 017202 (2012).
  • Iyer et al. (2013) S. Iyer, V. Oganesyan, G. Refael,  and D. A. Huse, Phys. Rev. B 87, 134202 (2013).
  • Huse et al. (2013) D. A. Huse, R. Nandkishore, V. Oganesyan, A. Pal,  and S. L. Sondhi, Phys. Rev. B 88, 014206 (2013).
  • Serbyn et al. (2013) M. Serbyn, Z. Papić,  and D. A. Abanin, Phys. Rev. Lett. 111, 127201 (2013).
  • Huse et al. (2014) D. A. Huse, R. Nandkishore,  and V. Oganesyan, Phys. Rev. B 90, 174202 (2014).
  • Kjäll et al. (2014) J. A. Kjäll, J. H. Bardarson,  and F. Pollmann, Phys. Rev. Lett. 113, 107204 (2014).
  • Nandkishore and Huse (2015) R. Nandkishore and D. A. Huse, Annual Review of Condensed Matter Physics 6, 15 (2015).
  • Bera et al. (2015) S. Bera, H. Schomerus, F. Heidrich-Meisner,  and J. H. Bardarson, Phys. Rev. Lett. 115, 046603 (2015).
  • Ponte et al. (2015) P. Ponte, Z. Papić, F. Huveneers,  and D. A. Abanin, Phys. Rev. Lett. 114, 140401 (2015).
  • Potter et al. (2015) A. C. Potter, R. Vasseur,  and S. A. Parameswaran, Phys. Rev. X 5, 031033 (2015).
  • Schreiber et al. (2015) M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Lüschen, M. H. Fischer, R. Vosk, E. Altman, U. Schneider,  and I. Bloch, Science 349, 842 (2015).
  • Luitz et al. (2015) D. J. Luitz, N. Laflorencie,  and F. Alet, Physical Review B 91, 081103(R) (2015).
  • Vosk et al. (2015) R. Vosk, D. A. Huse,  and E. Altman, Phys. Rev. X 5, 031032 (2015).
  • Smith et al. (2016) J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess, P. Hauke, M. Heyl, D. A. Huse,  and C. Monroe, Nature Physics 12, 907 (2016).
  • Choi et al. (2016) J.-y. Choi, S. Hild, J. Zeiher, P. Schauß, A. Rubio-Abadal, T. Yefsah, V. Khemani, D. A. Huse, I. Bloch,  and C. Gross, Science 352, 1547 (2016).
  • Imbrie (2016a) J. Z. Imbrie, Phys. Rev. Lett. 117, 027201 (2016a).
  • Imbrie (2016b) J. Z. Imbrie, Journal of Statistical Physics 163, 998 (2016b).
  • Khemani et al. (2017) V. Khemani, S. P. Lim, D. N. Sheng,  and D. A. Huse, Phys. Rev. X 7, 021013 (2017).
  • Alet and Laflorencie (2018) F. Alet and N. Laflorencie, Comptes Rendus Physique 19, 498 (2018).
  • Lukin et al. (2019) A. Lukin, M. Rispoli, R. Schittko, M. E. Tai, A. M. Kaufman, S. Choi, V. Khemani, J. Léonard,  and M. Greiner, Science 364, 256 (2019).
  • Abanin et al. (2019) D. A. Abanin, E. Altman, I. Bloch,  and M. Serbyn, Rev. Mod. Phys. 91, 021001 (2019).
  • Syassen et al. (2008) N. Syassen, D. M. Bauer, M. Lettner, T. Volz, D. Dietze, J. J. Garcia-Ripoll, J. I. Cirac, G. Rempe,  and S. Dürr, Science 320, 1329 (2008).
  • Barreiro et al. (2011) J. T. Barreiro, M. Müller, P. Schindler, D. Nigg, T. Monz, M. Chwalla, M. Hennrich, C. F. Roos, P. Zoller,  and R. Blatt, Nature 470, 486 (2011).
  • Yan et al. (2013) B. Yan, S. A. Moses, B. Gadway, J. P. Covey, K. R. Hazzard, A. M. Rey, D. S. Jin,  and J. Ye, Nature 501, 521 (2013).
  • Barontini et al. (2013) G. Barontini, R. Labouvie, F. Stubenrauch, A. Vogler, V. Guarrera,  and H. Ott, Phys. Rev. Lett. 110, 035302 (2013).
  • Labouvie et al. (2015) R. Labouvie, B. Santra, S. Heun, S. Wimberger,  and H. Ott, Phys. Rev. Lett. 115, 050601 (2015).
  • Patil et al. (2015) Y. S. Patil, S. Chakram,  and M. Vengalattore, Phys. Rev. Lett. 115, 140402 (2015).
  • Gao et al. (2015) T. Gao, E. Estrecho, K. Bliokh, T. Liew, M. Fraser, S. Brodbeck, M. Kamp, C. Schneider, S. Höfling, Y. Yamamoto, F. Nori, Y. Kivshar, A. Truscott, R. Dall,  and E. Ostrovskaya, Nature 526, 554 (2015).
  • Labouvie et al. (2016) R. Labouvie, B. Santra, S. Heun,  and H. Ott, Phys. Rev. Lett. 116, 235302 (2016).
  • Lüschen et al. (2017) H. P. Lüschen, P. Bordia, S. S. Hodgman, M. Schreiber, S. Sarkar, A. J. Daley, M. H. Fischer, E. Altman, I. Bloch,  and U. Schneider, Phys. Rev. X 7, 011034 (2017).
  • Raftery et al. (2014) J. Raftery, D. Sadri, S. Schmidt, H. E. Türeci,  and A. A. Houck, Phys. Rev. X 4, 031043 (2014).
  • Tomita et al. (2017) T. Tomita, S. Nakajima, I. Danshita, Y. Takasu,  and Y. Takahashi, Science advances 3, e1701513 (2017).
  • Lapp et al. (2019) S. Lapp, J. Ang’ong’a, F. A. An,  and B. Gadway, New Journal of Physics 21, 045006 (2019).
  • Takasu et al. (2020) Y. Takasu, T. Yagami, Y. Ashida, R. Hamazaki, Y. Kuno,  and Y. Takahashi, Progress of Theoretical and Experimental Physics 2020, 12A110 (2020).
  • Bouganne et al. (2020) R. Bouganne, M. B. Aguilera, A. Ghermaoui, J. Beugnon,  and F. Gerbier, Nature Physics 16, 21 (2020).
  • Keßler et al. (2021) H. Keßler, P. Kongkhambut, C. Georges, L. Mathey, J. G. Cosme,  and A. Hemmerich, Phys. Rev. Lett. 127, 043602 (2021).
  • Noel et al. (2021) C. Noel, P. Niroula, A. Risinger, L. Egan, D. Biswas, M. Cetina, A. V. Gorshkov, M. Gullans, D. A. Huse,  and C. Monroe, arXiv preprint arXiv:2106.05881  (2021).
  • Nandkishore et al. (2014) R. Nandkishore, S. Gopalakrishnan,  and D. A. Huse, Phys. Rev. B 90, 064203 (2014).
  • Johri et al. (2014) S. Johri, R. Nandkishore,  and R. Bhatt, arXiv preprint arXiv:1405.5515  (2014).
  • Fischer et al. (2016) M. H. Fischer, M. Maksymenko,  and E. Altman, Phys. Rev. Lett. 116, 160401 (2016).
  • Levi et al. (2016) E. Levi, M. Heyl, I. Lesanovsky,  and J. P. Garrahan, Phys. Rev. Lett. 116, 237203 (2016).
  • Medvedyeva et al. (2016) M. V. Medvedyeva, T. Prosen,  and M. Znidaric, Phys. Rev. B 93, 094205 (2016).
  • van Nieuwenburg et al. (2017) E. P. van Nieuwenburg, J. Y. Malo, A. J. Daley,  and M. H. Fischer, Quantum Science and Technology 3, 01LT02 (2017).
  • Morningstar et al. (2022) A. Morningstar, L. Colmenarez, V. Khemani, D. J. Luitz,  and D. A. Huse, Phys. Rev. B 105, 174205 (2022).
  • Sels (2021) D. Sels, arXiv preprint arXiv:2108.10796  (2021).
  • (73) References Morningstar et al. (2022); Sels (2021) analyzed the Liouvillian gap of the Lindbladian super-operator, which is different from the middle of the spectrum of the Lindbladian considered in our work as an analogue with the Hermitian case.
  • Hamazaki et al. (2019) R. Hamazaki, K. Kawabata,  and M. Ueda, Phys. Rev. Lett. 123, 090603 (2019).
  • Lindblad (1976) G. Lindblad, Comm. Math. Phys. 48, 119 (1976).
  • (76) In addition to the exponential decay whose rate is proportional to γ\gamma, the coherence may have a polynomial decay factor due to interactions as in Hermitian cases Serbyn and Abanin (2017); Abanin et al. (2019). However, this factor is negligible compared with the exponential decay.
  • (77) See Supplemental Material for details of the relaxation dynamics, structures of the overlaps concerning eigenstates, a phenomenology of the Lindbladian MBL phase, and dissipation dependence of the Lindbladian MBL transition point.
  • Gong and Hamazaki (2022) Z. Gong and R. Hamazaki, arXiv preprint arXiv:2202.02011  (2022).
  • (79) We normalize the eigenstates such that Tr[R^αR^α]=Tr[L^αL^α]=1\mathrm{Tr}[\hat{R}_{\alpha}^{\dagger}\hat{R}_{\alpha}]=\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{L}_{\alpha}]=1 is satisfied except for the stationary states, for which Tr[R^ss]=1\mathrm{Tr}[\hat{R}_{\mathrm{ss}}]=1 and L^ss=𝕀^\hat{L}_{\mathrm{ss}}=\hat{\mathbb{I}}.
  • (80) We note that a similar conjecture is given for classical Markovian systems, as discussed in Supplemental Material Sup .
  • Ginibre (1965) J. Ginibre, Journal of Mathematical Physics 6, 440 (1965).
  • Chalker and Mehlig (1998) J. T. Chalker and B. Mehlig, Phys. Rev. Lett. 81, 3367 (1998).
  • Mehlig and Chalker (2000) B. Mehlig and J. T. Chalker, Journal of Mathematical Physics 41, 3233 (2000).
  • Bourgade and Dubach (2019) P. Bourgade and G. Dubach, Probability Theory and Related Fields , 1 (2019).
  • (85) While the analytical formula for ensembles different from a complex Ginibre ensemble (e.g., a real Ginibre ensemble and an ensemble with transposition symmetry Hamazaki et al. (2020)) is not known, we have numerically confirmed this distribution. If λα\lambda_{\alpha} is real, then rαGr_{\alpha}^{G} and rαGr^{G^{\prime}}_{\alpha} can be taken to be real, and their distributions may be different from the case of complex eigenvalues. We here do not consider real eigenvalues because their fractions vanish in the limit of large system size.
  • (86) The peaks at x=0x=0 for Tr[L^αρ^]\mathrm{Tr}[\hat{L}_{\alpha}^{\dagger}\hat{\rho}] and Tr[𝒪^R^α]\mathrm{Tr}[\hat{\mathcal{O}}\hat{R}_{\alpha}] with large hh indicate that the eigenstates of the Lindbladian and those of ρ^\hat{\rho} or 𝒪^\hat{\mathcal{O}} can be orthogonal in the operator space for the localized regime.
  • (87) Precisely speaking, this result is for the coherence measured by, e.g., τ^lx\hat{\tau}_{l}^{x} or its product for different ll’s. While the dynamics of σ^lx\hat{\sigma}_{l}^{x} or its product may be slightly different from that for τ^lx\hat{\tau}_{l}^{x}, its difference becomes smaller if hh is increased. See Supplemental Material Sup for further discussions and additional numerical results.
  • Buča and Prosen (2012) B. Buča and T. Prosen, New Journal of Physics 14, 073007 (2012).
  • Albert and Jiang (2014) V. V. Albert and L. Jiang, Phys. Rev. A 89, 022118 (2014).
  • Šuntajs et al. (2020) J. Šuntajs, J. Bonča, T. Prosen,  and L. Vidmar, Physical Review E 102, 062144 (2020).
  • (91) Specifically, let us consider the real part λαr\lambda_{\alpha}^{r} and the absolute value |λαi||\lambda_{\alpha}^{i}| of the imaginary part of λα\lambda_{\alpha}, and sort each of them independently as a1aD2=0a_{1}\leq\cdots\leq a_{D^{2}}=0 and 0b1bD20\leq b_{1}\leq\cdots\leq b_{D^{2}}. We use only those modes whose eigenvalues satisfy a3D2/8λαra5D2/8a_{3D^{2}/8}\leq\lambda_{\alpha}^{r}\leq a_{5D^{2}/8} and b3D2/8|λαi|b5D2/8b_{3D^{2}/8}\leq|\lambda_{\alpha}^{i}|\leq b_{5D^{2}/8}.
  • Grobe et al. (1988) R. Grobe, F. Haake,  and H.-J. Sommers, Phys. Rev. Lett. 61, 1899 (1988).
  • Grobe and Haake (1989) R. Grobe and F. Haake, Phys. Rev. Lett. 62, 2893 (1989).
  • Markum et al. (1999) H. Markum, R. Pullirsch,  and T. Wettig, Phys. Rev. Lett. 83, 484 (1999).
  • Akemann et al. (2019) G. Akemann, M. Kieburg, A. Mielke,  and T. c. v. Prosen, Phys. Rev. Lett. 123, 254101 (2019).
  • Hamazaki et al. (2020) R. Hamazaki, K. Kawabata, N. Kura,  and M. Ueda, Phys. Rev. Research 2, 023286 (2020).
  • Jaiswal et al. (2019) A. B. Jaiswal, A. Pandey,  and R. Prakash, EPL (Europhysics Letters) 127, 30004 (2019).
  • Wang et al. (2020) K. Wang, F. Piazza,  and D. J. Luitz, Phys. Rev. Lett. 124, 100604 (2020).
  • Sá et al. (2020a) L. Sá, P. Ribeiro,  and T. c. v. Prosen, Phys. Rev. X 10, 021019 (2020a).
  • Sá et al. (2020b) L. Sá, P. Ribeiro, T. Can,  and T. c. v. Prosen, Phys. Rev. B 102, 134310 (2020b).
  • Tzortzakakis et al. (2020) A. F. Tzortzakakis, K. G. Makris,  and E. N. Economou, Phys. Rev. B 101, 014202 (2020).
  • Huang and Shklovskii (2020) Y. Huang and B. I. Shklovskii, Phys. Rev. B 101, 014204 (2020).
  • Mudute-Ndumbe and Graefe (2020) S. Mudute-Ndumbe and E.-M. Graefe, New Journal of Physics 22, 103011 (2020).
  • Luo et al. (2021) X. Luo, T. Ohtsuki,  and R. Shindou, Phys. Rev. Lett. 126, 090402 (2021).
  • Sá et al. (2021) L. Sá, P. Ribeiro,  and T. c. v. Prosen, Phys. Rev. B 103, 115132 (2021).
  • Rubio-García et al. (2022) A. Rubio-García, R. Molina,  and J. Dukelsky, SciPost Physics Core 5, 026 (2022).
  • Prasad et al. (2022) M. Prasad, H. K. Yadalam, C. Aron,  and M. Kulkarni, Phys. Rev. A 105, L050201 (2022).
  • García-García et al. (2022) A. M. García-García, L. Sá,  and J. J. M. Verbaarschot, Phys. Rev. X 12, 021040 (2022).
  • (109) For too strong hh, rr further decreases and becomes smaller than rPor_{\mathrm{Po}}.
  • Yusipov and Ivanchenko (2021) I. Yusipov and M. Ivanchenko, arXiv preprint arXiv:2112.06214  (2021).
  • Prosen and Pižorn (2007) T. c. v. Prosen and I. Pižorn, Phys. Rev. A 76, 032316 (2007).
  • Pižorn and Prosen (2009) I. Pižorn and T. c. v. Prosen, Phys. Rev. B 79, 184416 (2009).
  • (113) To define the OSEE, we consider a vector representation of R^α\hat{R}_{\alpha}, i.e., |Rα)=ij(R^α)ij|i|j|R_{\alpha})=\sum_{ij}(\hat{R}_{\alpha})_{ij}\ket{i}\otimes\ket{j}, where (Rα|Rα)=1({R_{\alpha}|R_{\alpha}})=1 because of the normalization condition. The entire system with the enlarged Hilbert space ~=\tilde{\mathcal{H}}=\mathcal{H}\otimes\mathcal{H}^{\prime} spanned by {|i|j||i,|j}{\left\{\ket{i}\otimes\ket{j}|\ket{i}\in\mathcal{H},\ket{j}\in\mathcal{H^{\prime}}\right\}} can be separated into two parts as ~=~AA~AA¯\tilde{\mathcal{H}}=\tilde{\mathcal{H}}_{AA}\otimes\tilde{\mathcal{H}}_{\overline{AA}}, where ~AA=AA\tilde{\mathcal{H}}_{AA}=\mathcal{H}_{A}\otimes\mathcal{H}_{A}^{\prime} and ~AA¯=A¯A¯\tilde{\mathcal{H}}_{\overline{AA}}=\mathcal{H}_{\overline{A}}\otimes\mathcal{H}_{\overline{A}}^{\prime} with A¯\overline{A} being the complement of the region AA. Using the reduced density matrix 𝚁^r=Tr~AA¯[|Rα)(Rα|]\hat{\mathtt{R}}_{r}=\mathrm{Tr}_{\tilde{\mathcal{H}}_{\overline{AA}}}[|{R_{\alpha}})({R_{\alpha}}|], we define the OSEE as Sα=Tr[𝚁^rln𝚁^r]S_{\alpha}=-\mathrm{Tr}[\hat{\mathtt{R}}_{r}\ln\hat{\mathtt{R}}_{r}].
  • Diehl et al. (2008) S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. Büchler,  and P. Zoller, Nature Physics 4, 878 (2008).
  • Yusipov et al. (2017) I. Yusipov, T. Laptyeva, S. Denisov,  and M. Ivanchenko, Phys. Rev. Lett. 118, 070402 (2017).
  • Vakulchyk et al. (2018) I. Vakulchyk, I. Yusipov, M. Ivanchenko, S. Flach,  and S. Denisov, Phys. Rev. B 98, 020202 (2018).
  • Haga et al. (2021) T. Haga, M. Nakagawa, R. Hamazaki,  and M. Ueda, Phys. Rev. Lett. 127, 070402 (2021).
  • Weinberg and Bukov (2017) P. Weinberg and M. Bukov, SciPost Phys. 2, 003 (2017).
  • Weinberg and Bukov (2019) P. Weinberg and M. Bukov, SciPost Phys 7, 97 (2019).
  • Serbyn and Abanin (2017) M. Serbyn and D. A. Abanin, Phys. Rev. B 96, 014202 (2017).