Metastable Random Liouvillian

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Random matrix theory for quantum and classical metastability in local Liouvillians

Jimin L. Li,1 Dominic C. Rose,2, 3, 4 Juan P. Garrahan,2, 3 and David J. Luitz5


1
Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge, CB2 1EW, UK
2
School of Physics and Astronomy, University of Nottingham, Nottingham, NG7 2RD, UK
3
Centre for the Mathematics and Theoretical Physics of Quantum Non-Equilibrium Systems,
University of Nottingham, Nottingham, NG7 2RD, UK
4
Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK
5
Max Planck Institute for the Physics of Complex Systems, Noethnitzer Str. 38, 01187 Dresden, Germany
(Dated: April 11, 2022)
We consider the effects of strong dissipation in quantum systems with a notion of locality, which
induces a hierarchy of many-body relaxation timescales as shown in [Phys. Rev. Lett. 124, 100604
(2020)]. If the strength of the dissipation varies strongly in the system, additional separations of
timescales can emerge, inducing a manifold of metastable states, to which observables relax first,
before relaxing to the steady state. Our simple model, involving one or two “good” qubits with
dissipation reduced by a factor α < 1 compared to the other “bad” qubits, confirms this picture
and admits a perturbative treatment.

Introduction — Quantum many-body systems are Here we apply such a local random matrix model ap-
generically complex, and obtaining an analytic under- proach to systems with strongly varying dissipation. We
standing of the position of all spectral resonances is often are specifically interested in the appearance of metastable
hopeless. It was realized early on [1–6] that this complex- states due to a separation of timescales caused by fast and
ity is in fact so great that many statistical properties of slow dissipation modes in the system, which we model by
the spectrum are identical with those of random matrices the existence of good qubits with low dissipation rates in
sampled from an ensemble determined by the symmetry a system of otherwise bad qubits where dissipation is fast.
of the system. These pioneering observations have been In this setup a metastable manifold (MM) emerges [46],
subsequently refined, resulting in cornerstones of our un- to which the dynamics starting from an arbitrary initial
derstanding of thermalization in unitary quantum many- state relaxes quickly. At intermediate times, the dynam-
body systems by virtue of the eigenstate thermalization ics is effectively restricted to the MM, before eventual
hypothesis [7–14], only with exceptions in integrable [15– relaxation to the global steady state at long times. We
18], many-body localized [19–29], time-crystalline [30–32] argue that this model contains the essence of the physics
or scarred and constrained systems [33–35]. to be expected in a quantum computer with good and
This thinking was recently pushed to the realm of bad qubits and is furthermore the simplest generic model
open quantum systems, with random matrix models of to study metastability. Our model generalizes findings of
Markovian dissipation defined via random Liouvillians MMs in the presence of local loss terms [47, 48].
[36–40], revealing fascinating spectral features of generic Model — We construct a simple model for a purely
purely dissipative systems, in particular a spectral sup- dissipative, Markovian quantum many-body system con-
port which has the shape of a “lemon” [36, 37], much dif- sisting of ℓ qubits. The Hilbert space dimension is
ferent from the circular spectrum of non-Hermitian Gini- N = 2ℓ , and the operator space is spanned by all N 2 = 4ℓ
bre random matrices [41]. This feature is also present in normalized Pauli strings
classical master equations, where typical transition rate
Sµ = N −1/2 σµ1 ⊗ σµ2 ⊗ · · · ⊗ σµℓ , µi ∈ {0, x, y, z}, (1)
matrices have a similar spectral support [42, 43].
Such random matrix models of open quantum many- where σ0 = 1, and σx,y,z are the Pauli matrices. Dis-
body systems represent the behavior of typical systems, sipation is generated by a set of k-local jump operators
rather than of a specific model. While they reproduce given by k-local Pauli strings such that the number of
global properties of more realistic, microscopic models, non-identity Pauli matrices in P the string is at most k.
l
they miss a crucial ingredient: the locality of (dissipative) That is, for k-local Sµ we have i=1 (1 − δµi ,0 ) ≤ k. We
interactions. It was recently shown that random matrix will focus on the physically relevant case of two-body dis-
models for local Liouvillians can be devised exhibiting sipative interactions, including one qubit and two qubit
a hierarchy of relaxation timescales [44]. These models (k = 2) dissipation channels, yielding NL = 3ℓ + 9 2ℓ
limit the jump operators in the Lindblad equation to low jump operators.
complexity Pauli strings, thus encoding few-body interac- The dynamics of the density matrix ρ is governed by
tions. In the absence of detailed microscopic knowledge, the purely dissipative Liouvillian [49] defined in terms of
this accurately models dissipation in current quantum a Kossakowski matrix Kµν which encodes the nontrivial
computer prototypes, and the predicted timescales were couplings between the dissipation channels and is ran-
in fact observed experimentally on the IBM platform [45]. domly sampled from the ensemble of positive semidefinite
2

matrices. The sums µ, ν run over the NL jump operators α = 0.5 α = 0.5
0.02 gap= 0.2094 gap= 0.2069
Lµ = Sµ , given by k-local Pauli strings in Eq. (1),

Im(λ)
NL   0.00
X
† 1 †
L[ρ] = Kµν Lµ ρLν − {Lν Lµ , ρ} . (2)
µ,ν=1
2
−0.02
# good qubits = 1 # good qubits = 2
Using the same procedure as in Refs. [36, 44, 45], we α = 0.1 α = 0.1
generate the i.i.d. non-negative eigenvalues of K from 0.02 gap= 0.0421 gap= 0.0415

a uniform distribution, and normalize them such that

Im(λ)
TrK = N . Then, we rotate the basis by a Haar random 0.00

unitary U ∈ CUE(NL ) to yield K = U † DU , where D is


the diagonal eigenvalue matrix of K. −0.02
# good qubits = 1 # good qubits = 2
In contrast to Ref. [44], we are interested in un-
α = 0.01 α = 0.01
derstanding the effect of a strongly varying dissipation 0.02 gap= 0.0042 gap= 0.0042
strength across the system. The simplest way to consider

Im(λ)
this is by splitting the set of jump operators {Lµ } into 0.00
strongly dissipative ones, {Lsµ }, and weakly dissipative
ones, {Lwµ }. This is achieved by defining a set of “good” −0.02
# good qubits = 1 # good qubits = 2
qubits in a system of otherwise “bad” qubits: weak jump
operators are those that contain a non-identity Pauli ma- −0.5 0.0 −0.5 0.0
Re(λ) Re(λ)
trix on a good√qubit, so that √ dissipation happens at a
rate scaled by α < 1, Lw µ = αSµ , while strong jump FIG. 1. Spectra λi of random local Liouvillians for ℓ = 6 with
operators are still of the form Lsµ = Sµ [50]. one (left column) or two (right column) good qubits as a func-
Spectrum of the Liouvillian — In Fig. 1, we show tion of the weak dissipation rate α. The spectral gap, given by
the complex eigenvalues λi of a realization of the Li- the magnitude of the real part of the first excited eigenvalue,
ouvillian for ℓ = 6, for one (left) and two (right) good decreases proportionally with α. Bars indicate the eigenval-
ues of the unperturbed Liouvillian L0 , the starting point of
qubits, with two qubit interactions and one qubit dissi-
our perturbation theory. Blue bars indicate the position of
pation (k = 2-local in our definition). The Liouvillian (2) eigenvalues giving rise to the MM at small α. For small α
is bi-stochastic (as all Lµ are Hermitian), thus it generi- there is a separation between metastable eigenmodes and the
cally has a single eigenvalue zero with the identity as the rest of the spectrum, since all other eigenvalues are have real
unique stationary state, and all other eigenvalues with parts smaller than λ(ns = 1, nw = 0) given in Eq. (4), and
negative real parts. indicated by the vertical dashed line in each panel.
Due to the locality of our model, the spectrum sepa-
rates into multiple eigenvalue clusters, organized by the
locality of their eigenmodes. If good and bad qubits have √ 
3/2
standard deviation is N/ 6NL , which can be shown
the same rate of dissipation (α = 1), we recover the
spectrum of Ref. [44]. As we make good qubits better by the central limit theorem and using random matrix
(α < 1), additional eigenvalue clusters appear. These theory for U . Hence, we can devise a perturbative treat-
clusters have a real part proportional to α, and are indi- ment by splitting the K matrix as K = K0 + K1 , where
cated by the blue bars in Fig. 1. For decreasing α, these K0 = (N/NL ) 1 is the unperturbed matrix, and K1 a
clusters move progressively closer to zero. For small α small perturbation which we neglect for now.
they combine to form the MM (see below) of long lived We can express the Liouvillian L (or it adjoint) as a
states with the slowest relaxation. The other clusters also matrix in the Pauli string basis [51] with matrix elements
move slightly with α, but reach a limiting position well L0µν = Tr (Sµ L [Sν ]). To leading order off-diagonal ma-
separated from the MM. trix elements vanish, and by separating the expressions
Note that in the case of one good qubit, there is a for weak and strong channels we get for the diagonal el-
single cluster of three eigenvalues close to zero, while for ements and thus eigenvalues to leading order,
two good qubits there are two such clusters, one with six
and the other with nine eigenvalues. To elucidate these
2
L0µµ = − αNµw + Nµs ,

spectral properties further, we study these clusters using (3)
perturbation theory. NL
Perturbation Theory — Due to the physical re-
quirement that the Liouvillian is trace preserving and where Nµw and Nµs are the numbers of weak and strong
completely positive, the random K matrix is diagonally jump operators, respectively, that anticommute with Sµ .
dominant [44]. It hence has the following properties: the The number nw and ns of non-identity Pauli matrices
mean of the matrix elements Kµν is δµ,ν N/NL and the on good and bad qubits determine the above. For the
3

2-local case with ℓw good qubits, we obtain from Eq. (3) where we have split the contribution of the M eigenvalues
with largest real parts from the rest of the spectrum, and
2 h were the coefficients cm are given by cm = Tr(Lm ρ0 ), Lm
λ(ns , nw ) = − 6ns ℓ + 6ns ℓw (α − 1)
NL (4) being the left eigenmatrices. For a large spectral separa-
i
− 4n2s + α(6nw ℓ − 8nw ns − 4n2w ) . tion, there is a wide range of times for which the modes
m > M have already decayed and can be neglected above
Since there are many Pauli strings with the same num- giving
bers of nw and ns , each eigenvalue is highly degenerate. M
X
There is a unique steady state, nw = ns = 0, correspond- ρ(t) ≈ R0 + eλm t cm Rm . (6)
ing to the identity. m=1
Including the small perturbation K1 lifts the degener-
acy of the eigenvalues, and gives them small imaginary This is the metastable regime where dynamics is approxi-
parts. To see this, we diagonalize the Liouvillian with mately restricted to the lower dimensional MM. The valid
K = K1 inside each degenerate subspace. Lowly de- combinations of ci classify a MM as either classical or
generate eigenvalues, which are well separated from the quantum [46, 55]. A MM is called classical if there exists
rest of the spectrum, develop into the clusters observed in a basis of density matrices ρ̃i so that any state Pmin the
Fig. 1, while for eigenvalues close to others and with high MM is a positive linear combination ρ(t) ≈ i=1 pi ρ̃i
degeneracy the separation does not survive, and the per- with 0 ≤ pi ≤ 1. In this case the MM is a simplex, anal-
turbation theory breaks down in these cases (see [52] for ogous to the manifold of probability distributions, with
a detailed discussion of this for large systems). For small the pi the probabilities of being in each metastable phase
α, the perturbation theory is excellent and yields well ρ̃i , and the long time dynamics can be cast as a classical
separated eigenvalue clusters close to the steady state, as Markov jump process between these phases. When such
can be seen in Fig. 1. Each eigenvalue cluster in this case basis does not exist the MM is said to be quantum.
is centered around the unperturbed eigenvalue λ(ns , nw ), At the level of the perturbative calculation above, we
indicated by black (ns > 0) and blue (ns = 0) bars, as can read off the eigenmatrices Rm and Lm (m ≤ M )
predicted from our perturbation theory. forming the MM. For one good qubit (ℓw = 1), we have
Further inspection of Eq. (4) reveals that eigenval- three eigenvalues with nw = 1, ns = 0, and the matri-
ues corresponding to observables with only identities on ces are the three Pauli strings with a non-identity on the
bad qubits (i.e. ns = 0) are proportional to −α, while good qubit and the identity. For two good qubits, we
any observable with a non-identity on a bad qubit picks obtain two eigenvalue clusters in the MM, which remain
up a constant offset and thus generically has a much well separated even for large ℓ (cf. discussion in [44] and
faster decay rate. This is what makes up the MM: eigen- [52]): one is formed by the six one qubit Pauli strings
values with real parts proportional to −α are close to with one identity on one of the good qubits (nw = 1),
zero for small α, and well separated from the rest of the and the other by the nine two qubit Pauli strings with
spectrum. They are nonidentities on both good qubits (nw = 2). Perturba-
4α 2
 centered around λ(ns = 0, nw ) = tion theory thus suggests that the MM is quantum, since
−N L
3n w ℓ − 2n w , which means that for one good qubit
we get one eigenvalue cluster (since nw can only be ei- it is invariant under the action of SU(2) operators on
ther zero or one), and for ℓw good qubits, we get ℓw sepa- the slow sites. However, this assumption might fail when
rate eigenvalue clusters with eigenvalues proportional to we take into account the full random K matrix, the ad-
−α. The remaining eigenvalues are always smaller than ditional corrections allowing for a classical manifold to
λ(ns = 1, nw = 0) = − N4L (3(ℓ − ℓw ) − 2 + 3ℓw α) , indi- form. We now test this numerically.
cated by the vertical dashed line in Fig. 1. This sets the To test for classicality of the MM we apply the algo-
separation between the MM and the rest of the spectrum, rithmic approach of Refs. [54, 55] which tries to system-
and thus the relaxation timescale of an arbitrary initial atically find the best possible simplex from spectral data
state towards the MM before relaxation to the steady of the Liouvillian. Accuracy is measured by a bound on
state happens on a timescale ∝ 1/α. the average distance of metastable states outside this op-
Metastable manifold — The existence of eigenval- timal simplex. This bound follows by noting that given
ues with small real parts, which are well separated from some basis ρ̃m (m = 0, . . . , M ) of the MM [56]. They
the rest of the spectrum for small α gives rise to metasta- correspond to the ”metastable phases” that coexist dur-
bility [46, 53–55]. The evolution ρ(t) = etL ρ0 of any ini- ing the long transient regime of metastability, in , there
tial state ρ0 can be written in terms of the eigenvalues exists a unique dual basis P̃m with normalization chosen
λm and right eigenmatrices Rm of the Liouvillian, as Tr(P̃m ρ̃m′ ) = δmm′ . Therefore, the coefficients for a
state ρ projected to the MM are given by pm = Tr(P̃m ρ),
M
X 4
X
ℓ bounded by the maximum and minimum eigenvalues of
ρ(t) = R0 + eλm t cm Rm + eλm t cm Rm , (5) P̃m . These eigenvalues reside between 0 and 1 if the
m=1 m=M +1 MM is exactly a simplex, and thus classical. How far
4

(P )
any eigenvalues λj m of the P̃m are outside of this range
defines a classicality measure [55]
M 2N
1 XX h
(P )
i
C= N max −λj m , 0 . (7)
2 m=0 j=1

Since an exactly classical MM has vanishing C, the more


C departs from zero the further away from classical the
MM is.
With this procedure we construct the simplex approx-
imation to the MM for a set of 1000 realizations of the
disorder matrix K at a dissipation rate α = 10−5 , show-
ing a histogram of C in Fig. 2(a) (green). We see that
the manifold is never classical, C ≳ 1, in the disorder
realizations we consider. To illustrate this visually, for
one sample realization we plot the projections of random
pure states on to the metastable manifold against the
expectation values of Pauli operators on the slow site in
Fig. 2(c): we see that many of these projections fall out-
side the optimal simplex, and indication that the MM is
not classical. In Fig. 2(e), we evolve a few metastable
states as they converge towards the stationary state, see-
ing that some spend time outside the simplex (but still
within the quantum MM).
The quantum nature is apparently robust in this ran-
FIG. 2. (a) Histograms of the classicality C for an ensem-
dom matrix model as suggested by perturbation theory. ble of 1000 disorder realizations of K, for the quantum MM
It is natural to expect that the non-commuting dissipa- case (green, right) and the classical MM case (purple, left),
tion channels are responsible for the robustness of the for l = 6 and at a dissipation rate α = 10−5 . (b) Simplex
quantum MM, e.g., the long time dynamics for ℓw = 1 that best approximates the MM in the classical MM case
lives on a Bloch sphere, which a classical master equa- for one disorder realization (blue dots indicate the extreme
tion cannot approximate. To construct a classical MM, metastable phases [56], lines the edges of the simplex, dashed
we manually remove this feature by making only the z lines are behind the volume of the simplex) in which approxi-
mately all metastable states are contained if the MM is classi-
direction of the good
√ qubits strongly dissipative, i.e., the cal, in the basis of Pauli strings with Z Pauli matrices on the
dissipation rate α = 1 for the Z Pauli matrices on good slow sites 1 and 6 [S1z = Sz00000 , S6z = S0000z , S1z S6z = Sz0000z
qubits. For example, we consider this modified model for in the notation of Eq. (1)]. Dots (purple) are projections of
ℓw = 2 and {Lw µ } are those jump operators that have X a set of random initial states onto the MM, plotted accord-
or Y Pauli matrices on the good qubits, but not those ing to the expectation value of the three observables. All
with Z. Counterintuitively, the MM is spanned by the states sampled fall within the simplex, as expected for a clas-
strongly dissipative Z Pauli strings on the good qubits sical MM. (c) Same for the quantum MM case, now in terms
of Pauli strings with a non-trivial Pauli matrix on the slow
because such strings commute with all rapidly relaxing site [S2x = S0x0000 , S2y = S0y0000 , S2z = S0z0000 ]. Projections
operators in Eq. (3). To obtain a classical MM, we slow of random initial states escape the simplex, as the MM is
only certain Pauli operators
√ on the good qubits. For ex- quantum (the shaded Bloch sphere), with those inside colored
ample, we multiply by α only those jump operators that green and those outside colored red. (d,e) Projections of the
have X or Y Pauli matrices on the ℓw = 2 good qubits, time-evolution for long times of the metastable phases (blue
but not those with Z. In the perturbation theory, this curves) and of a set of random initial states (black curves)
results in only the Z operators on good qubits commut- within the MM towards stationarity (red dot), for the classi-
cal MM case (left panel) and quantum MM case (right panel).
ing with all rapidly relaxing operators in Eq. (3). In this Curve segments colored red in (e) are outside the simplex ap-
case, the MM is thus made up of 4 operators: the iden- proximation.
tity, the Z operator on each good qubit, and the product
of Z operators on both good qubits. The algorithms of
[54, 55] yield an extremely accurate simplex approxima-
tion to the MM, confirming that it is effectively classical set of these metastable states (black), or the metastable
as shown in Fig. 2(a). This is visualized by projecting phases (blue) remain within the simplex at all times.
a set of random states onto the slow-mode eigenspace in Metastable dynamics — Using Eq. (5), we can
Fig. 2(b), locating them well within the simplex. Fur- calculate the evolution of observables at any time. To
ther, as shown in Fig. 2(d), the long-time evolution of a consider generic initial states we choose ρ0 as a random
5

1.00 α = 10−2 α = 10−2


can lead to classical manifolds instead.
0.75
This work was supported in part by the Deutsche
Tr(ρ0 O(t))

one body
0.50 two body Forschungsgemeinschaft through SFB 1143 (Project-id
0.25
three body 247310070) and the cluster of excellence ML4Q (EXC
2004). JPG and DCR acknowledge financial support
0.00 from EPSRC Grant no. EP/R04421X/1 and the Lev-
# good qubits= 1 # good qubits= 2
1.00 α = 10 −5
α = 10−5 erhulme Trust Grant No. RPG-2018-181.
0.75
Tr(ρ0 O(t))

0.50

0.25
[1] E. P. Wigner, Characteristic Vectors of Bordered Matri-
0.00 ces With Infinite Dimensions, Ann. Math. 62, 548 (1955).
# good qubits= 1 # good qubits= 2
100 103 106 100 103 106 [2] E. P. Wigner, Random Matrices in Physics, SIAM Rev.
t t 9, 1 (1967).
[3] F. J. Dyson, Statistical Theory of the Energy Levels of
Complex Systems. I, J. Math. Phys. 3, 140 (1962).
FIG. 3. Comparison of the exact dynamics (solid lines) with [4] F. J. Dyson, Statistical Theory of the Energy Levels of
the dynamics projected onto the MM (dashed lines) for ℓ = 7, Complex Systems. II, J. Math. Phys. 3, 157 (1962).
for qubits with one (left column) and two good (right column) [5] F. J. Dyson, Statistical Theory of the Energy Levels of
qubits at two different dissipation rates α = 10−2 , 10−5 . We Complex Systems. III, J. Math. Phys. 3, 166 (1962).
show expectation values of three observables with different lo- [6] M. L. Mehta, Random Matrices (Elsevier, 2004).
calities, each prepared by a random linear superposition of all [7] M. Feingold, N. Moiseyev, and A. Peres, Ergodicity and
the Pauli matrices of the corresponding locality. Observables mixing in quantum theory. II, Phys. Rev. A 30, 509
supported by the bad qubits vanish rapidly, and the long-time (1984).
dynamics on the good qubits coincide with the effective dy- [8] J. M. Deutsch, Quantum statistical mechanics in a closed
namics. Note that for a MM to display, the locality of the system, Phys. Rev. A 43, 2046 (1991).
observable has to be smaller or equal to the number of the [9] J. M. Deutsch, Eigenstate thermalization hypothesis,
good qubits. Rep. Prog. Phys. 81, 082001 (2018).
[10] M. Srednicki, Chaos and quantum thermalization, Phys.
Rev. E 50, 888 (1994).
linear superposition of the full Hilbert space. Figure 3 [11] M. Srednicki, Thermal fluctuations in quantized chaotic
(full lines) shows the time evolution of observables with systems, J. Phys. A: Math. Gen. 29, L75 (1996).
[12] M. Rigol, V. Dunjko, and M. Olshanii, Thermalization
different locality properties (non-trivial Pauli strings of
and its mechanism for generic isolated quantum systems,
different lengths k). Note the appearance of plateaus in Nature 452, 854 (2008).
the relaxation curves, specifically for the shorter Pauli [13] L. D’Alessio, Y. Kafri, A. Polkovnikov, and M. Rigol,
strings which have a larger overlap with the matrices Rm From Quantum Chaos and Eigenstate Thermalization to
that define the MM. Statistical Mechanics and Thermodynamics, Adv. Phys.
After a fast transient, dynamics is confined to the MM. 65, 239 (2016).
The approximate dynamics is then obtained by project- [14] F. Borgonovi, F. M. Izrailev, L. F. Santos, and V. G.
Zelevinsky, Quantum chaos and thermalization in iso-
ing both the initial state and the observables onto it, and
lated systems of interacting particles, Phys. Rep. 626,
solving Eq. (6). The dashed curves in Fig. 3 show the cor- 1 (2016).
responding results: the effective dynamics captures the [15] M. V. Berry, M. Tabor, and J. M. Ziman, Level clustering
long-time behavior accurately, showing that metastabil- in the regular spectrum, Proc. R. Soc. A 356, 375 (1977).
ity implies dimensional reduction from the whole Hilbert [16] T. Kinoshita, T. Wenger, and D. S. Weiss, A quantum
space to the MM. Newton’s cradle, Nature 440, 900 (2006).
Conclusion — Starting from a random local and [17] M. Rigol, V. Dunjko, V. Yurovsky, and M. Olshanii, Re-
laxation in a Completely Integrable Many-Body Quan-
purely dissipative Liouvillian, we have defined a ran- tum System: An Ab Initio Study of the Dynamics of the
dom matrix model for generic metastability in open Highly Excited States of 1D Lattice Hard-Core Bosons,
quantum systems relevant for strongly varying dissipa- Phys. Rev. Lett. 98, 050405 (2007).
tion timescales in quantum computers if dissipation is [18] T. Langen, S. Erne, R. Geiger, B. Rauer, T. Schweigler,
strongly dominant. We find that a separation of dissi- M. Kuhnert, W. Rohringer, I. E. Mazets, T. Gasenzer,
pation timescales induces the presence of a metastable and J. Schmiedmayer, Experimental observation of a gen-
manifold to which initial states relax, before the evolu- eralized Gibbs ensemble, Science (2015).
[19] P. W. Anderson, Absence of Diffusion in Certain Random
tion to the steady state occurs at much longer times. If Lattices, Phys. Rev. 109, 1492 (1958).
the dissipation on good qubits does not further single [20] D. M. Basko, I. L. Aleiner, and B. L. Altshuler, Metal–
out certain Pauli operators, we show that the metastable insulator transition in a weakly interacting many-electron
manifold is generically quantum, while further structure system with localized single-particle states, Ann. Phys.
6

321, 1126 (2006). Spectral gaps and mid-gap states in random quantum
[21] V. Oganesyan and D. A. Huse, Localization of interacting master equations, arXiv:1902.01414 (2019).
fermions at high temperature, Phys. Rev. B 75, 155111 [40] S. Lange and C. Timm, Random-matrix theory for the
(2007). Lindblad master equation, Chaos 31, 023101 (2021).
[22] M. Žnidaric, T. Prosen, and P. Prelovšek, Many-body [41] J. Ginibre, Statistical Ensembles of Complex, Quater-
localization in the Heisenberg XXZ magnet in a random nion, and Real Matrices, J. Math. Phys. 6, 440 (1965).
field, Phys. Rev. B 77, 064426 (2008). [42] C. Timm, Random transition-rate matrices for the mas-
[23] A. Pal and D. A. Huse, Many-body localization phase ter equation, Phys. Rev. E 80, 021140 (2009).
transition, Phys. Rev. B 82, 174411 (2010). [43] W. Tarnowski, I. Yusipov, T. Laptyeva, S. Denisov,
[24] D. J. Luitz, N. Laflorencie, and F. Alet, Many-body local- D. Chruściński, and K. Życzkowski, Random generators
ization edge in the random-field Heisenberg chain, Phys. of Markovian evolution: A quantum-classical transition
Rev. B 91, 081103(R) (2015). by superdecoherence, Phys. Rev. E 104, 034118 (2021).
[25] R. Nandkishore and D. A. Huse, Many-Body Localization [44] K. Wang, F. Piazza, and D. J. Luitz, Hierarchy of Re-
and Thermalization in Quantum Statistical Mechanics, laxation Timescales in Local Random Liouvillians, Phys.
Annu. Rev. Condens. Matter Phys. 6, 15 (2015). Rev. Lett. 124, 100604 (2020).
[26] D. A. Abanin, E. Altman, I. Bloch, and M. Serbyn, Col- [45] O. E. Sommer, F. Piazza, and D. J. Luitz, Many-body hi-
loquium: Many-body localization, thermalization, and erarchy of dissipative timescales in a quantum computer,
entanglement, Rev. Mod. Phys. 91, 021001 (2019). Phys. Rev. Research 3, 023190 (2021).
[27] D. J. Luitz and Y. Bar Lev, Anomalous Thermalization [46] K. Macieszczak, M. Guţă, I. Lesanovsky, and J. P. Gar-
in Ergodic Systems, Phys. Rev. Lett. 117, 170404 (2016). rahan, Towards a Theory of Metastability in Open Quan-
[28] D. J. Luitz and Y. Bar Lev, The ergodic side of the many- tum Dynamics, Phys. Rev. Lett. 116, 240404 (2016).
body localization transition, Ann. Phys. 529, 1600350 [47] H. Fröml, A. Chiocchetta, C. Kollath, and S. Diehl,
(2017). Fluctuation-Induced Quantum Zeno Effect, Phys. Rev.
[29] D. A. Abanin and Z. Papić, Recent progress in many- Lett. 122, 040402 (2019).
body localization, Ann. Phys. 529, 1700169 (2017). [48] S. Wolff, A. Sheikhan, S. Diehl, and C. Kollath, Nonequi-
[30] V. Khemani, A. Lazarides, R. Moessner, and librium metastable state in a chain of interacting spinless
S. L. Sondhi, Phase Structure of Driven Quantum fermions with localized loss, Phys. Rev. B 101, 075139
Systems, Phys. Rev. Lett. 116, 250401 (2016). (2020).
[31] D. V. Else, B. Bauer, and C. Nayak, Floquet Time Crys- [49] C. Gardiner and P. Zoller, Quantum noise (Springer,
tals, Phys. Rev. Lett. 117, 090402 (2016). 2004).
[32] V. Khemani, R. Moessner, and S. L. Sondhi, A Brief [50] Note that we do not impose a notion of geometry, essen-
History of Time Crystals, arXiv:1910.10745 (2019). tially dealing with a zero dimensional system. A geometry
[33] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Om- can be added on top as discussed in the supplement of
ran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres, Ref. [44], and adds further complexity, which we do not
M. Greiner, V. Vuletić, and M. D. Lukin, Probing many- discuss here to focus on the essential physics of metasta-
body dynamics on a 51-atom quantum simulator, Nature bility.
551, 579 (2017). [51] Note that the normalization is automatic since we defined
[34] C. J. Turner, A. A. Michailidis, D. A. Abanin, M. Serbyn, Pauli strings such that TrSµ2 = 1.
and Z. Papić, Weak ergodicity breaking from quantum [52] Supplemental Material.
many-body scars, Nature Phys. 14, 745 (2018). [53] D. C. Rose, K. Macieszczak, I. Lesanovsky, and J. P. Gar-
[35] N. Pancotti, G. Giudice, J. I. Cirac, J. P. Garrahan, and rahan, Metastability in an open quantum ising model,
M. C. Bañuls, Quantum east model: Localization, non- Phys. Rev. E 94, 052132 (2016).
thermal eigenstates, and slow dynamics, Phys. Rev. X [54] D. C. Rose, K. Macieszczak, I. Lesanovsky, and J. P.
10, 021051 (2020). Garrahan, Hierarchical classical metastability in an open
[36] S. Denisov, T. Laptyeva, W. Tarnowski, D. Chruściński, quantum east model (2020), arXiv:2010.15304.
and K. Życzkowski, Universal Spectra of Random Lind- [55] K. Macieszczak, D. C. Rose, I. Lesanovsky, and J. P. Gar-
blad Operators, Phys. Rev. Lett. 123, 140403 (2019). rahan, Theory of classical metastability in open quantum
[37] L. Sá, P. Ribeiro, and T. Prosen, Spectral and systems, Phys. Rev. Research 3, 033047 (2021).
Steady-State Properties of Random Liouvillians, [56] The states ρ̃m are called ‘extreme metastable states’ [55].
arXiv:1905.02155 (2019). They define the endpoints of the simplex that best ap-
[38] T. Can, Random Lindblad Dynamics, arXiv:1902.01442 proximates a classical MM, and all metastable states are
(2019). probabilistic mixtures of them.
[39] T. Can, V. Oganesyan, D. Orgad, and S. Gopalakrishnan,

You might also like