You are on page 1of 16

Observation of Time-Crystalline Eigenstate Order on a Quantum Processor

Google Quantum AI and collaborators†,∗

Quantum many-body systems display rich ෡


𝐻(𝑡) ෡0
= 𝐻 ෡ + 𝑇) = 𝐻(𝑡)
𝐻(𝑡 ෡
≃ |01001⟩
phase structure in their low-temperature equilib- a b c +|10110⟩

rium states1 . However, much of nature is not


Ordered
in thermal equilibrium. Remarkably, it was re-

Energy
cently predicted that out-of-equilibrium systems Unordered

can exhibit novel dynamical phases2–8 that may


≃ |00000⟩ − |11111⟩
≃ |01001⟩
otherwise be forbidden by equilibrium thermo- ≃ |00000⟩ + |11111⟩ −|10110⟩
dynamics, a paradigmatic example being the dis-
arXiv:2107.13571v2 [quant-ph] 11 Aug 2021

crete time crystal (DTC)7,9–14 . Concretely, dy- FIG. 1. Order in eigenstates. a, Equilibrium phases
namical phases can be defined in periodically are characterized by long-range order in low-energy eigen-
driven many-body localized systems via the con- states of time-independent Hamiltonians, e.g. an Ising fer-
cept of eigenstate order7,15,16 . In eigenstate- romagnet with a pair of degenerate ground states that resem-
ordered phases, the entire many-body spectrum ble “Schrödinger cats” of polarized states. b, Floquet sys-
exhibits quantum correlations and long-range or- tems typically have no ordered states in the spectrum. c, In
der, with characteristic signatures in late-time MBL Floquet systems, every eigenstate can show order. In
dynamics from all initial states. It is, how- MBL-DTC, every eigenstate resembles a long-range ordered
ever, challenging to experimentally distinguish “Schrödinger cat” of a random configuration of spins and its
such stable phases from transient phenomena, inversion, with even/odd superpositions split by π.
wherein few select states can mask typical be-
havior. Here we implement a continuous fam-
ily of tunable CPHASE gates on an array of su- circle. While the entire Floquet spectrum is featureless
perconducting qubits to experimentally observe in a thermalizing phase (Fig. 1b), an MBL Floquet phase
an eigenstate-ordered DTC. We demonstrate the can have an order parameter associated with each eigen-
characteristic spatiotemporal response of a DTC state. As an example, in the spatiotemporally-ordered
for generic initial states 7,9,10 . Our work employs a MBL-DTC, the spectrum has a distinctive pattern of
time-reversal protocol that discriminates external pairing between “Schrödinger cat” states that are sep-
decoherence from intrinsic thermalization, and arated by an angle π (Fig. 1c)7,9,10 . This pairing man-
leverages quantum typicality to circumvent the ifests as a stable sub-harmonic response, wherein local
exponential cost of densely sampling the eigen- observables show period-doubled oscillations that sponta-
spectrum. In addition, we locate the phase tran- neously break the discrete time translation symmetry of
sition out of the DTC with an experimental finite- the drive for infinitely long times. The unique combina-
size analysis. These results establish a scalable tion of spatial long-range order and time translation sym-
approach to study non-equilibrium phases of mat- metry breaking in an isolated dissipation-free quantum
ter on current quantum processors. many-body system is the hallmark of the MBL-DTC.
In an equilibrium setting, quantum phases of mat- Experimentally observing a non-equilibrium phase
ter are classified by long-range order or broken sym- such as the MBL-DTC is a challenge due to limited
metries in low-temperature states (Fig. 1a). The ex- programmability, coherence and size of Noisy Interme-
istence of ordered phases in periodically driven (Flo- diate Scale Quantum (NISQ) hardware. Sub-harmonic
quet) systems, on the other hand, is counterintuitive: response, by itself, is not a unique attribute of the MBL-
Since energy is not conserved, one expects thermaliza- DTC; rather, it is a feature of many dynamical phe-
tion to a featureless maximum-entropy state that is in- nomena whose study has a rich history23,24 (See also
compatible with quantum order. However, this heat Ch. 8 in Ref.12 ). Most recently, interesting DTC-like
death is averted in the presence of many-body localiza- dynamical signatures have been observed in a range
tion (MBL), where strong disorder causes the emergence of quantum platforms25–28 . Such signatures, however,
of an extensive number of local conservation laws which are transient, arising from slow or prethermal dynam-
prevent thermalization17–22 , making it possible to stabi- ics from special initial states12,29–32 , and are separated
lize intrinsically dynamical phases7 . from the MBL-DTC by a spectral phase transition where
Dynamics in a Floquet system is governed by a unitary eigenstate order disappears. Thus, despite the recent
time evolution operator, whose eigenvalues lie on the unit progress, observing an MBL-DTC remains an outstand-
ing challenge12,32 .
Here we perform the following necessary benchmarks
∗ Corresponding authors: V. Khemani: vkhemani@stanford.edu; for experimentally establishing an eigenstate-ordered
P. Roushan: pedramr@google.com non-equilibrium phase of matter: (i) Drive parameters
2

a Random Bit-strings Random i and hi ෡ c Disorder and Initial-State Averaged Autocorrelators


Z(t)
Xg Z (h1) 1.0
Q1 : 0 1 0 ഥ ෡Ft ) ഥ (𝑈
෡Ft )

ZZ (12)
A (𝑈 A
Xg Z (h2)
ഥ 0 (𝑈
A ෡ECHO) ഥ ෡ECHO)
A0 (𝑈
Q2 : 1 0 1

ZZ (23)
0.5

෡ Z(t)
Q3 : 0 0 1 Xg Z (h3)

ZZ (34)


0.0

Z(0)
Xg Z (h4)
෡ECHO
𝑈
Q4 : 1 0 0

ZZ (45)
-0.5
Q5 : 1 1 0 Xg Z (h5) ෡Ft
𝑈 ෡F† )t
(𝑈
g = 0.60, Qubit 11 g = 0.97, Qubit 11
1 Xg Z (h20) -1.0
Q20: 0 0
0 10 20 0 20 40 60 80 100
෡F (t Cycles)
𝑈 t t

b Thermal MBL-DTC d Thermal (g = 0.60) MBL-DTC (g = 0.97)


1.0 1.0
g = 0.60, g = 0.97,
Qubit 11 Qubit 11 0.5

ഥA ഥ0
0.5 0.0

A/
-0.5
-1.0 ഥA
A/ ഥ0
Z(t)

0.0

20 1.0
Instance 1 15 0.5
Qubit
-0.5 Instance 2
Instance 3 10 0.0
Instance 4 5 -0.5
Instance 5 -1.0
-1.0
0 10 20 0 20 40 60 80 100 0 10 20 0 20 40 60 80 100
t t t t

FIG. 2. Observing a many-body localized discrete time-crystal. a, Schematic of the experimental circuit composed
of t identical cycles of the unitary ÛF . The local polarization of each qubit, hẐ(t)i, is measured at the end. In the following
panels, we investigate a number of disorder instances each with a different random bit-string initial state. b, Experimental
values of hẐ(t)i measured at qubit 11. Data are shown for five representative circuit instances deep in the thermal (g = 0.60)
and MBL-DTC (g = 0.97) phases. c, Autocorrelator A = hẐ(0)Ẑ(t)i at qubit 11, obtained from averaging the results of 36
circuit instances. For the same circuit instances, the average autocorrelator at the output of ÛECHO = (ÛF† )t ÛFt is also measured
and its square root, A0 , is shown alongside A for comparison. d, Top panels: The ratio A/A0 obtained from panel c. Bottom
panels: A/A0 as a function of t and qubit location.

are varied in order to demonstrate stability of the phase explore the DTC phase and its transition into a thermal
in an extended parameter region and across disorder real- phase. At g = 1, the model implements a π pulse which
izations; The limitations of (ii) finite size and (iii) finite exactly flips all qubits (in the z basis) and returns them
coherence time are addressed, respectively, by varying to the initial state over two periods. A key signature of
the number of qubits in the system and by separating the DTC is the presence of robust period doubling, i.e.
effects of extrinsic decoherence from intrinsic thermaliza- extending over a finite extent in parameter space, even
tion; (iv) The existence of eigenstate order across the en- as g is tuned away from 1. Strong Ising interactions,
tire spectrum is established. The flexibility of our quan- which produce long-range spatial order, are essential for
tum processor, combined with the scalable experimental this robustness7,10 . This is in contrast to a system of
protocols devised in the following, allows us to fulfill these decoupled qubits (φ = 0) which rotate by a continuously
criteria and observe time-crystalline eigenstate order. varying angle πg every period instead of being locked
The experiment is conducted on an open-ended, linear at period doubling. Prior theoretical work32 has shown
chain of L = 20 superconducting transmon qubits (Q1 that model (1) is expected to be in an MBL DTC phase
through Q20 ) that are isolated from a two-dimensional in the range g > gc , and transition to a thermal phase at
grid. We drive the qubits via a time-periodic (Floquet) a critical value gc ≈ 0.84.
circuit ÛFt with t identical cycles (Fig. 2a) of ÛF : Achieving MBL in this model for g ∼ 1 requires disor-
−2 i
P
−4 i
P
− 2 πgi
P der in the two-qubit Qinteraction, φi , which is even under
ÛF = i hi Ẑi φi Ẑi Ẑi+1 X̂i
|e {z } |e } e| (1) Ising symmetry12,32 , i X̂i , a condition that was not met
i i
{z {z }
longitudinal fields Ising interaction x rotation by πg by some past DTC experiments26,27 . Ising-odd terms,
i.e. hi , are approximately dynamically decoupled by the
where X̂i and Ẑi are Pauli operators. Each φi (hi ) is x pulses over two periods, thereby lowering their effective
sampled randomly from [−1.5π, −0.5π] ([−π, π]) for ev- disorder strength and hindering localization (in the ab-
ery realization of the circuit. Overall, ÛF implements an sence of independent disorder in the φi ). Utilizing newly
interacting Ising model that is periodically “kicked” by a developed CPHASE gates (see SM for details) with con-
transverse pulse that rotates all qubits by πg about the x tinuously tunable conditional phases allows us to engi-
axis. In this work, g is tuned within the range [0.5, 1.0] to neer strong disorder in φi to fulfill this key requirement.
3

a c Initial State 1: 00000000000000000000 Initial State 2: 00000000001000000000


1.0 Qubit 11 Qubit 12 Qubit 13 Qubit 14
i  [-1.5, -0.5] i = -0.4
0.5 0.5

Z(t)
0.0
[A]

0.0


-0.5 Néel Néel
-0.5 i  [-1.5, -0.5]
Polarized Polarized
Random Random 0.5
-1.0
0 20 40 60 0 20 40 60 ζ2

Z(t)
t t 0.0


ζ1 |ζ1−ζ2|
i = -0.4 i  [-1.5, -0.5] ζr =
b -0.5 |ζ1|+|ζ2| i = -0.4
30 40 50 60 30 40 50 60 30 40 50 60 30 40 50 60
σ/μ 0.4 σ/μ t
30 = 0.129 = 0.038 i  [-1.5, -0.5]
d 0.0 0.2 0.4 0.6 0.8 1.0
ζതr i = -0.4
Number of States

0.3
|[A]|

20 1.0
i  [-1.5, -0.5] i = -0.4
20 0.2
15 0.8
0.1

Qubit
0.6
-2 -1 0 1 2

ζതr
10
10 Approximate Energy (a.u) 0.4
5 0.2

0 0.0
0.10 0.15 0.20 0.25 0.30 0.35 0.40 0 20 40 60 80 100 0 20 40 60 80 100 5 10 15 20

|[A]| t t Qubit

FIG. 3. Observing eigenstate order and distinguishing it from transient phenomena. a, Site- and disorder-averaged
autocorrelators [A] measured with g = 0.94. In the left panel (MBL-DTC), each data set is averaged over 24 disorder instances of
φi and hi , with the initial state fixed at one of the following: Néel: |01i⊗10 , Polarized: |0i⊗20 , Random: |00111000010011001111i.
In the right panel (thermalizing), the same values of hi and initial states are used but φi = −0.4. b, Histograms of |[A]|, from
500 random bit-string initial states, averaged over cycles 30 and 31 and the same disorder instances as in panel a. σ/µ, where σ
(µ) is the standard deviation (mean) of |[A]|, is also listed. Location of the polarized (Néel) state is indicated by a purple (red)
arrow. Inset: same collection of |[A]| plotted over the energies of the bit-string states, calculated from the effective Hamiltonian
Ĥeff approximating the drive (see text). Dashed lines show averaged values within energy windows separated by 0.2. c, hẐ(t)i
for two bit-string initial states that differ only at Q11 . Top panel shows a single circuit instance with disordered φi and bottom
panel shows an instance with uniform φi = −0.4. d, Left and middle panels: Relative difference between the two signals ζ r
as a function of t and qubit location, averaged over time windows of 10 cycles and over 64 disorder instances for ÛF and 81
instances for ÛF0 . Right panel: Qubit dependence of ζ r , averaged from t = 51 to t = 60.

We first measure the hallmark of an MBL-DTC: the ÛECHO from the identity operation are purely due to de-
persistent oscillation of local qubit polarizations hẐ(t)i coherence, and can be quantified via decay of the auto-

at a period twice that of ÛF , irrespective of the initial correlator A0 ≡ (hẐ ÛECHO Ẑ ÛECHO i)1/2 (the square root
state7,9,12,32 . This subharmonic response is probed using accounts for the fact that ÛECHO acts twice as long as
a collection of random bit-string states, e.g. |01011...i ÛFt ). A similar time-reversal technique was recently used
where 0(1) denotes a single-qubit ground(excited) state in the study of out-of-time-ordered commutators in ther-
in the z basis. For each bit-string state, we generate a malizing random circuits33 .
random instance of ÛF , and then measure hẐ(t)i every Comparison between the disorder averaged A0 and A
cycle. Figure 2b shows hẐ(t)i in a few different instances reveals qualitatively different behaviors in the two phases
for a qubit near the center of the chain, Q11 , measured (Fig. 2c). In the thermal phase g = 0.60, A approaches
with g = 0.60 and g = 0.97. The former is deep in 0 much more quickly than A0 does, indicating that the
the thermal phase and indeed we observe rapid decay of observed decay of A is mostly induced by intrinsic ther-
hẐ(t)i toward 0 within 10 cycles for each instance. In malization. In the MBL-DTC phase g = 0.97, A0 nearly
contrast, for g = 0.97, hẐ(t)i shows large period-doubled coincides with the envelope of A, suggesting that decay
oscillations persisting to over 100 cycles, suggestive of an of the latter is primarily induced by decoherence. The
MBL-DTC phase. The disorder averaged autocorrelator, reference signal A0 may be used to normalize A and re-
A = hẐ(0)Ẑ(t)i, shows similar features (Fig. 2c). veal its ideal behavior: A/A0 , shown in the upper panels
of Fig. 2d, decays rapidly for g = 0.60 but retains near-
We note that the data for g = 0.97 is modulated by a maximal amplitudes for g = 0.97. Similar contrast be-
gradually decaying envelope, which may arise from either tween the two phases is seen in the error-mitigated auto-
external decoherence or slow internal thermalization25,30 . correlators A/A0 for all qubits (bottom panel of Fig. 2d).
To establish DTC, additional measurements are needed The observation of a stable noise-corrected sub-harmonic
to distinguish between these two mechanisms. This is response is suggestive of an MBL-DTC phase.
achieved via an “echo” circuit ÛECHO = (ÛF† )t ÛFt which We now turn to a systematic analysis of the next re-
reverses the time evolution after t steps. Deviations of quirement necessary to establish an MBL-DTC: namely
4

a ෡ Z(t)|ψ
ψ|Z(0) ෡ ෡
=X
b |ψ〉: Product state Random State affected by the presence of edge modes independent of
a

Qa : |+Xۧ
K=0 K~L the bulk DTC response36 ). The three time traces are
|ψ〉 Cycle 1 Cycle 2 Cycle K nearly indistinguishable. This behavior is in clear con-
R1,2 R1,K
Q1
R1,1
trast with a model without eigenstate order, implemented
by a family of drives ÛF0 where the φi angles are set to a
|Bit−stringۧ

R2,1 R2,2 R2,K


Q2

Q11
෡S
𝑈 ෡Ft
𝑈 Random Single-Qubit Rotations
uniform value37 , φi = −0.4. Without disorder in the φi ,
CZ CZ
R20,1 R20,2 R20,K
the drive ÛF0 is not asymptotically localized but exhibits
Q20
transient DTC-like behavior. Here, [A] for ÛF0 (disorder
c d
1.0
Qubit 11, g = 0.94
averaged over random hi only), shown in the right panel
40 σ/μ = K=0
of Fig 3a, reveals markedly different decay rates for the
෡ Z(t)|ψ

0.5
0.015 K=2
20 three states. The random bit-string state, in particular,

0.0
ψ|Z(0)

K = 20

Number of States
-0.5 Aψ (𝑈෡Ft ) 0 decays faster than the polarized or Néel states.
-1.0
Aψ,0 (𝑈෡ECHO) 40 A more comprehensive analysis, presented in Fig. 3b,
0.031
1.0 20 is based on sampling the absolute values of [A] for 500
0.5 0 random initial bit-string states (averaged over cycles 30
Aψ/Aψ,0

0.0 40 0.062
and 31). For the MBL-DTC ÛF , the histogram is tight,
-0.5 20 with a relative standard deviation (ratio of standard de-
-1.0 0
viation to mean, σ/µ) of 0.038. Here the non-zero value
0 20 40
t
60 80 100 0.35 0.40
|Aψ|
0.45 of σ likely stems from finite experimental accuracy and
number of disorder instances, as analysis in the SM shows
that [A] is independent of the initial state. In contrast,
FIG. 4. Probing average spectral response via quan-
tum typicality. a, Schematic for measuring the auto- the ÛF0 model shows a broader distribution with a much
correlator, Aψ = hψ| Ẑ(0)Ẑ(t) |ψi, on Q11 , of a scrambled lower mean, and with σ/µ = 0.129. Moreover, the his-
quantum state |ψi. |ψi is created by passing a bit-string state togram is asymmetrical, with outliers at high [A] includ-
through a scrambling circuit, ÛS . An ancilla qubit Qa pre- ing the polarized and Néel states (51% and 88% higher
pared in |+Xi interacts with one of the qubits (Q11 ) via a CZ than the mean, respectively). These two states are spe-
gate before and after ÛFt . The x-axis projection of Qa , hX̂ia , cial because they are low temperature states that sit near
is measured at the end. b, ÛS contains K layers of CZ gates the edge of the spectrum of Ĥeff (see SM). Plotting the
interleaved with random single-qubit rotations, Ri,k , around autocorrelator [A] against the energy of each bitstring
a random axis along the equatorial plane of the Bloch sphere under Ĥeff , in the inset of Fig. 3b, reveals a clear corre-
by an angle ∈ [0.4π, 0.6π]. c, Upper panel: Aψ for a single lation. No such correlation is present in the MBL model.
disorder instance with K = 20 cycles in ÛS . The square-root Independent confirmation of MBL as the mechanism
of the autocorrelator obtained by replacing ÛFt with ÛECHO , underlying the stability of DTC is achieved by character-
Aψ,0 , is also shown. Bottom panel: Normalized autocorrela-
izing the propagation of correlations. In MBL dynamics,
tor, Aψ /Aψ,0 , as a function of t. d, Histograms of |Aψ | from
a single disorder instance, averaged over cycles 30 and 31.
local perturbations spread only logarithmically in time19 ,
Each histogram corresponds to a different number of scram- as opposed to algebraic (∼ tα ) spreading in thermalizing
bling cycles, K, and includes data from 500 random initial dynamics. We prepare two initial bitstring states differ-
bit-string states fed through the scrambling circuit. ing by only a single bit-flip at Q11 and measure hẐ(t)i
for each site in both states (Fig. 3c). It can be seen that
the difference in the two signals, ζ1 and ζ2 , decays rapidly
with the distance from Q11 for disordered φi and becomes
the presence of eigenstate order across the entire spec- undetectable at Q14 . On the other hand, for uniform
trum which, in turn, implies that sub-harmonic response φi = −0.4, ζ1 and ζ2 have a much more pronounced dif-
should not be strongly affected by the choice of initial ference which remains significant at Q14 . This difference
states. In contrast, various prethermal mechanisms in is further elucidated by the ratio ζr = |ζ1 −ζ2 |/(|ζ1 |+|ζ2 |),
driven systems predict strong dependence of the ther- shown in Fig. 3d. Physically, ζr corresponds to the rela-
malization rate on the initial state, e.g. through its quan- tive change in local polarization as a result of the bit flip,
tum numbers27,31 or its energy under an effective time- and is inherently robust against qubit decoherence (see
independent Hamiltonian Ĥeff 29,34,35 that approximately SM). We observe that up to t = 100, ζr remains sharply
governs the dynamics for small system sizes and/or finite peaked around the initial perturbation (Q11 ) for disor-
times. To elucidate this aspect of the MBL-DTC phase, dered φi . In contrast, a propagating light cone is visible
we analyze in detail the distribution of autocorrelator for φi = −0.4, with the perturbation reaching all qubits
values over initial bit-string states. across the chain as t increases. The spatial profiles of ζr
We begin by examining the position- and disorder- at t = 51 to t = 60 (right panel of Fig. 3d) show that
averaged autocorrelator [A] over three representative bit- ζr is much sharper for disordered φi . This slow prop-
string initial states, shown in the left panel of Fig 3a. agation provides strong indication of MBL and another
The square brackets indicate averaging over qubits in the experimental means of distinguishing eigenstate-ordered
chain (excluding the edge qubits Q1 , Q20 , which may be phases from transient phenomena.
5

Our measurement of [A] for 500 initial states in Fig. 3d 0.03


provides clear evidence of initial state independence. 0.030
g = 0.90

Still, a direct sampling of states is practically limited to 0.025

Spin Glass Order Parameter, SG


small fractions of the computational basis (0.05% in this
0.020
case) and would suffer from the exponential growth of

SG
0.02
g = 0.78
the Hilbert space on larger systems. A more scalable al- 0.0022
0.0020
ternative is to use random, highly entangled states to di- 0.0018
rectly measure spectrally-averaged quantities (quantum 0.0016
typicality38–40 ). The autocorrelator A averaged over all 8 12 16 20
0.01
2L bitstrings agrees, up to an error exponentially small Number of Qubits, L L=8
in L, with Aψ = hψ| Ẑ(0)Ẑ(t) |ψi, where |ψi is a typical L = 12
Haar-random many-body state in the Hilbert space of L L = 16
qubits. We prepare such a state by evolving a bitstring L = 20
0.00
with a random circuit ÛS of variable depth K (Fig. 4b),
0.72 0.74 0.76 0.78 0.80 0.82 0.84 0.86 0.88 0.90
and couple an ancilla qubit to the system to measure g
the two-time operator Ẑ(0)Ẑ(t) (Fig. 4a). Experimen-
tal results for the error-mitigated, spectrally averaged FIG. 5. Estimating phase-transition by varying system
signal Aψ /Aψ,0 on qubit Q11 (Fig. 4c) show behavior size. Spin-glass order parameter χSG as a function of g for dif-
consistent with a stable MBL-DTC. The effect of the ferent chain lengths L, measured between t = 51 and t = 60.
state-preparation circuit ÛS is illustrated by the depen- Every data point is averaged over 40 disorder instances. To
dence of the relative standard deviation σ/µ for Aψ on construct χSG , we sample 40000 bit-strings at the output of
K. As shown in Fig. 4d, σ/µ steadily decreases as K ÛFt for each cycle and disorder instance. To address the in-
increases, reducing from a value of 0.062 at K = 0 to a homogeneity of qubit coherence, smaller qubit chains are also
value of 0.015 at K = 20. This is consistent with the averaged over different possible combinations of qubits. For
fact that |ψi becomes closer to a Haar-random state as example, L = 12 is averaged over 12-qubit chains made from
Q1 through Q12 , Q3 through Q15 etc. Error bars correspond
K increases. We use a single disorder instance to study
to statistical errors (estimated by resampling data from the
the convergence of the quantum typicality protocol be- 40 disorder instances via the jackknife method). Contribution
cause disorder averaging independently leads to narrow of hardware (e.g. gate) errors to χSG is not included in the
distributions even for K = 0 (Fig. 3b). error bars. Inset shows the size-dependence of χSG for two
The scaling with L of the spectrally-averaged autocor- different values of g.
relator, at a time t ∼ poly(L), provides a sharp diag-
nostic: this saturates to a finite value in the MBL-DTC,
while it scales to zero with increasing L in transient cases random (“glassy”) spatial patterns in the initial bitstring
(where, for instance, a vanishing fraction of the spectrum state: at late times, it vanishes with increasing L in the
of an appropriate Ĥeff shows order). While the averaged thermalizing phase g < gc , while it is extensive in the
autocorrelator may be unduly affected by outlier states MBL-DTC g > gc . As a result, it is expected to show a
and/or long (but O(1)) thermalization times at small sys- finite-size crossing at g ' gc (though the precise location
tem sizes and times (thereby making the complementary is subject to strong finite-size and finite-time drifts44,45 ).
bitstring analysis of Fig. 3 essential), the polynomial scal- Experimentally, χSG is constructed from bit-string sam-
ing of this protocol establishes a proof of principle for ples obtained by jointly reading out all qubits and then
efficiently verifying the presence or absence of an MBL averaged over cycles and disorder instances (Fig. 5). The
DTC in a range of models as quantum processors scale size of the qubit chain is varied by restricting the drive
up in size to surpass the limits of classical simulation41 . ÛF to contiguous subsets of 8, 12, and 16 qubits (as well
Finally, we systematically vary g in small increments as the entire 20-qubit chain). We observe increasing (de-
and obtain an experimental finite-size analysis to estab- creasing) trends in χSG vs L when g is above (below) a
lish the extent of the MBL phase and the transition out critical value gc . The data indicate 0.83 . gc . 0.88,
of it. Sharply defining phases of matter, whether in or consistent with numerical simulations (see SM).
out of equilibrium, requires a limit of large system size.
In conclusion, we have demonstrated the possibility of
Thus it is important to examine the stability of the MBL-
engineering and characterizing non-equilibrium phases of
DTC and thermalizing regimes observed in our finite-size
matter on a quantum processor, providing direct experi-
quantum processor as the size of the system is increased.
mental observation of an eigenstate-ordered MBL-DTC.
To address this, we measure an Edwards-Anderson spin
The scalability of our protocols sets a blueprint for future
glass order parameter42,43 ,
studies of non-equilibrium phases and phase transitions
1 X0 2 on complex quantum systems beyond classical simulabil-
χSG = hẐi Ẑj i (2)
L−2 ity. The efficient verification of eigenstate order can in-
i6=j
spire a general strategy for establishing whether a desired
(the primed sum excludes edge qubits Q1 , QL ), as a func- property, such as a particular phase, is in fact present in
tion of time. This quantity measures the persistence of a quantum processor.
6


Google Quantum AI and Collaborators

Xiao Mi1,∗ , Matteo Ippoliti2,∗ , Chris Quintana1 , Ami Greene1, 3 , Zijun Chen1 , Jonathan Gross1 , Frank Arute1 , Kunal Arya1 ,
Juan Atalaya1 , Ryan Babbush1 , Joseph C. Bardin1, 4 , Joao Basso1 , Andreas Bengtsson1 , Alexander Bilmes1 , Alexandre
Bourassa1, 7 , Leon Brill1 , Michael Broughton1 , Bob B. Buckley1 , David A. Buell1 , Brian Burkett1 , Nicholas Bushnell1 ,
Benjamin Chiaro1 , Roberto Collins1 , William Courtney1 , Dripto Debroy1 , Sean Demura1 , Alan R. Derk1 , Andrew
Dunsworth1 , Daniel Eppens1 , Catherine Erickson1 , Edward Farhi1 , Austin G. Fowler1 , Brooks Foxen1 , Craig Gidney1 ,
Marissa Giustina1 , Matthew P. Harrigan1 , Sean D. Harrington1 , Jeremy Hilton1 , Alan Ho1 , Sabrina Hong1 , Trent Huang1 ,
Ashley Huff1 , William J. Huggins1 , L. B. Ioffe1 , Sergei V. Isakov1 , Justin Iveland1 , Evan Jeffrey1 , Zhang Jiang1 , Cody
Jones1 , Dvir Kafri1 , Tanuj Khattar1 , Seon Kim1 , Alexei Kitaev1 , Paul V. Klimov1 , Alexander N. Korotkov1, 6 , Fedor
Kostritsa1 , David Landhuis1 , Pavel Laptev1 , Joonho Lee1, 8 Kenny Lee1 , Aditya Locharla1 , Erik Lucero1 , Orion Martin1 ,
Jarrod R. McClean1 , Trevor McCourt1 , Matt McEwen1, 5 , Kevin C. Miao1 , Masoud Mohseni1 , Shirin Montazeri1 , Wojciech
Mruczkiewicz1 , Ofer Naaman1 , Matthew Neeley1 , Charles Neill1 , Michael Newman1 , Murphy Yuezhen Niu1 , Thomas
E. O’Brien1 , Alex Opremcak1 , Eric Ostby1 , Balint Pato1 , Andre Petukhov1 , Nicholas C. Rubin1 , Daniel Sank1 , Kevin
J. Satzinger1 , Vladimir Shvarts1 , Yuan Su1 , Doug Strain1 , Marco Szalay1 , Matthew D. Trevithick1 , Benjamin Villalonga1 ,
Theodore White1 , Z. Jamie Yao1 , Ping Yeh1 , Juhwan Yoo1 , Adam Zalcman1 , Hartmut Neven1 , Sergio Boixo1 , Vadim
Smelyanskiy1 , Anthony Megrant1 , Julian Kelly1 , Yu Chen1 , S. L. Sondhi9 , Roderich Moessner10 , Kostyantyn Kechedzhi1 ,
Vedika Khemani2 , Pedram Roushan1

1
Google Research, Mountain View, CA, USA
2
Department of Physics, Stanford University, Stanford, CA, USA
3
Department of Physics, Massachusetts Institute of Technology, Cambridge, MA, USA
4
Department of Electrical and Computer Engineering, University of Massachusetts, Amherst, MA, USA
5
Department of Physics, University of California, Santa Barbara, CA, USA
6
Department of Electrical and Computer Engineering, University of California, Riverside, CA, USA
7
Pritzker School of Molecular Engineering, University of Chicago, Chicago, IL, USA
8
Department of Chemistry, Columbia University, New York, NY 10027, USA
9
Department of Physics, Princeton University, Princeton, NJ, USA
10
Max-Planck-Institut für Physik komplexer Systeme, 01187 Dresden, Germany
∗ These authors contributed equally to this work.

Acknowledgements—This work was supported in part from the Defense Advanced Research Projects Agency (DARPA) via the
DRINQS program (M.I., V.K., R.M., S.S.), by a Google Research Award: Quantum Hardware For Scientific Research In Physics (V.K.
and M.I.), and from the Sloan Foundation through a Sloan Research Fellowship (V.K.) The views, opinions and/or findings expressed are
those of the authors and should not be interpreted as representing the official views or policies of the Department of Defense or the U.S.
Government. M.I. was funded in part by the Gordon and Betty Moore Foundation’s EPiQS Initiative through Grant GBMF8686. This
work was partly supported by the Deutsche Forschungsgemeinschaft under grants SFB 1143 (project-id 247310070) and the cluster of
excellence ct.qmat (EXC 2147, project-id 390858490).

Author contributions—M.I., K.K., V.K., R.M. and S.S. conceived the project. X.M., C.Q. and P.R. executed the experiment. All the
aforementioned discussed the project in progress and interpreted the results. M.I., K.K., V.K. and X.M., designed measurement
protocols. J.C., A.G., J.G. and X.M. implemented and calibrated the CPHASE gates. M.I. and V.K. performed theoretical and
numerical analyses. M.I., K.K., V.K., R.M., X.M., and P.R. wrote the manuscript. M.I. and X.M. wrote the supplementary material.
Y.C., K.K., V.K., H.N., P.R. and V.S. led and coordinated the project. Infrastructure support was provided by Google Quantum AI. All
authors contributed to revising the manuscript and the supplementary material.

Data availability—The data presented in this work are available from the corresponding authors upon reasonable request.

[1] Xiao-Gang Wen, Quantum field theory of many-body sys- graphene,” Physical Review B 79 (2009), 10.1103/phys-
tems: from the origin of sound to an origin of light and revb.79.081406.
electrons (Oxford University Press, Oxford, 2007). [5] Mark S. Rudner, Netanel H. Lindner, Erez Berg,
[2] Marin Bukov, Luca D’Alessio, and Anatoli Polkovnikov, and Michael Levin, “Anomalous edge states and the
“Universal high-frequency behavior of periodically driven bulk-edge correspondence for periodically driven two-
systems: from dynamical stabilization to floquet engi- dimensional systems,” Phys. Rev. X 3, 031005 (2013).
neering,” Advances in Physics 64, 139–226 (2015). [6] Paraj Titum, Erez Berg, Mark S. Rudner, Gil Refael,
[3] Fenner Harper, Rahul Roy, Mark S. Rudner, and S.L. and Netanel H. Lindner, “Anomalous floquet-anderson
Sondhi, “Topology and broken symmetry in floquet sys- insulator as a nonadiabatic quantized charge pump,”
tems,” Annual Review of Condensed Matter Physics 11, Phys. Rev. X 6, 021013 (2016).
345–368 (2020). [7] V. Khemani, A. Lazarides, R. Moessner, and S. Sondhi,
[4] Takashi Oka and Hideo Aoki, “Photovoltaic hall effect in “Phase structure of driven quantum systems,” Phys. Rev.
7

Lett. 116, 250401 (2016). von Keyserlingk, Norman Y. Yao, Eugene Demler,
[8] Hoi Chun Po, Lukasz Fidkowski, Takahiro Morimoto, and Mikhail D. Lukin, “Observation of discrete time-
Andrew C. Potter, and Ashvin Vishwanath, “Chiral flo- crystalline order in a disordered dipolar many-body sys-
quet phases of many-body localized bosons,” Phys. Rev. tem,” Nature 543, 221–225 (2017).
X 6, 041070 (2016). [26] J. Zhang, P. W. Hess, A. Kyprianidis, P. Becker, A. Lee,
[9] Dominic V. Else, Bela Bauer, and Chetan Nayak, “Flo- J. Smith, G. Pagano, I.-D. Potirniche, A. C. Potter,
quet time crystals,” Phys. Rev. Lett. 117, 090402 (2016). A. Vishwanath, N. Y. Yao, and C. Monroe, “Obser-
[10] C. W. von Keyserlingk, Vedika Khemani, and S. L. vation of a discrete time crystal,” Nature 543, 217–220
Sondhi, “Absolute stability and spatiotemporal long- (2017).
range order in floquet systems,” Phys. Rev. B 94, 085112 [27] Jared Rovny, Robert L. Blum, and Sean E. Barrett,
(2016). “Observation of discrete-time-crystal signatures in an or-
[11] Krzysztof Sacha and Jakub Zakrzewski, “Time crystals: dered dipolar many-body system,” Phys. Rev. Lett. 120,
a review,” Reports on Progress in Physics 81, 016401 180603 (2018).
(2017). [28] Soham Pal, Naveen Nishad, T. S. Mahesh, and G. J.
[12] V. Khemani, R. Moessner, and S. Sondhi, “A brief his- Sreejith, “Temporal order in periodically driven spins
tory of time crystals,” arXiv e-prints , arXiv:1910.10745 in star-shaped clusters,” Phys. Rev. Lett. 120, 180602
(2019). (2018).
[13] Patrick Bruno, “Impossibility of Spontaneously Rotating [29] D. Else, B. Bauer, and C. Nayak, “Prethermal phases of
Time Crystals: A No-Go Theorem,” Phys. Rev. Lett. matter protected by time-translation symmetry,” Phys.
111, 070402 (2013), arXiv:1306.6275 [quant-ph]. Rev. X 7, 011026 (2017).
[14] Haruki Watanabe and Masaki Oshikawa, “Absence of [30] Wen Wei Ho, Soonwon Choi, Mikhail D. Lukin, and
quantum time crystals,” Physical Review Letters 114 Dmitry A. Abanin, “Critical time crystals in dipolar sys-
(2015), 10.1103/physrevlett.114.251603. tems,” Phys. Rev. Lett. 119, 010602 (2017).
[15] David A. Huse, Rahul Nandkishore, Vadim Oganesyan, [31] David J. Luitz, Roderich Moessner, S. L. Sondhi, and
Arijeet Pal, and S. L. Sondhi, “Localization-protected Vedika Khemani, “Prethermalization without tempera-
quantum order,” Phys. Rev. B 88, 014206 (2013). ture,” Phys. Rev. X 10, 021046 (2020).
[16] David Pekker, Gil Refael, Ehud Altman, Eugene Demler, [32] Matteo Ippoliti, Kostyantyn Kechedzhi, Roderich Moess-
and Vadim Oganesyan, “Hilbert-glass transition: New ner, S. L. Sondhi, and Vedika Khemani, “Many-body
universality of temperature-tuned many-body dynamical physics in the NISQ era: quantum programming a dis-
quantum criticality,” Phys. Rev. X 4, 011052 (2014). crete time crystal,” arXiv e-prints , arXiv:2007.11602
[17] D.M. Basko, I.L. Aleiner, and B.L. Altshuler, “Metal- (2020), arXiv:2007.11602 [cond-mat.dis-nn].
insulator transition in a weakly interacting many-electron [33] X. Mi, P. Roushan, C. Quintana, S. Mandra, J. Mar-
system with localized single-particle states,” Annals of shall, C. Neill, F. Arute, K. Arya, J. Atalaya, R. Bab-
Physics 321, 1126–1205 (2006). bush, et al., “Information scrambling in computationally
[18] R. Nandkishore and D. A. Huse, “Many-body localiza- complex quantum circuits,” (2021), arXiv:2101.08870
tion and thermalization in quantum statistical mechan- [quant-ph].
ics,” Annu. Rev. Condens. Matter Phys. 6, 15–38 (2015). [34] D. A. Abanin, W. De Roeck, W. W. Ho, and F. Hu-
[19] Dmitry A. Abanin, Ehud Altman, Immanuel Bloch, and veneers, “Effective Hamiltonians, prethermalization, and
Maksym Serbyn, “Colloquium: Many-body localization, slow energy absorption in periodically driven many-body
thermalization, and entanglement,” Rev. Mod. Phys. 91, systems,” Phys. Rev. B 95, 014112 (2017).
021001 (2019). [35] Takashi Mori, Tatsuhiko N Ikeda, Eriko Kaminishi, and
[20] Pedro Ponte, Z. Papic, Francois Huveneers, and Masahito Ueda, “Thermalization and prethermalization
Dmitry A. Abanin, “Many-body localization in peri- in isolated quantum systems: a theoretical overview,”
odically driven systems,” Physical Review Letters 114 Journal of Physics B: Atomic, Molecular and Optical
(2015), 10.1103/physrevlett.114.140401. Physics 51, 112001 (2018).
[21] Achilleas Lazarides, Arnab Das, and Roderich Moess- [36] Daniel J. Yates, Alexander G. Abanov, and Aditi Mitra,
ner, “Fate of many-body localization under periodic driv- “Long-lived π edge modes of interacting and disorder-free
ing,” Physical Review Letters 115 (2015), 10.1103/phys- floquet spin chains,” (2021), arXiv:2105.13766 [cond-
revlett.115.030402. mat.str-el].
[22] Pranjal Bordia, Henrik Luschen, Ulrich Schneider, [37] This value of φi = −0.4 is chosen to be small enough
Michael Knap, and Immanuel Bloch, “Periodically driv- that a leading order high-frequency Floquet Magnus ex-
ing a many-body localized quantum system,” Nature pansion to obtain Ĥeff is a reasonable approximation (see
Physics 13, 460–464 (2017). SM).
[23] M. Faraday, “On a peculiar class of acoustical figures; [38] Sandu Popescu, Anthony J. Short, and Andreas Win-
and on certain forms assumed by groups of particles upon ter, “Entanglement and the foundations of statistical me-
vibrating elastic surfaces,” Philosophical Transactions of chanics,” Nature Physics 2, 754–758 (2006).
the Royal Society of London 121, 299–340 (1831). [39] Sheldon Goldstein, Joel L. Lebowitz, Roderich Tumulka,
[24] Raymond E. Goldstein, “Coffee stains, cell receptors, and Nino Zanghı̀, “Canonical typicality,” Phys. Rev.
and time crystals: Lessons from the old literature,” Lett. 96, 050403 (2006).
Physics Today 71, 32–38 (2018), arXiv:1811.08179 [cond- [40] Jonas Richter and Arijeet Pal, “Simulating hydrodynam-
mat.soft]. ics on noisy intermediate-scale quantum devices with ran-
[25] Soonwon Choi, Joonhee Choi, Renate Landig, Georg dom circuits,” Phys. Rev. Lett. 126, 230501 (2021).
Kucsko, Hengyun Zhou, Junichi Isoya, Fedor Jelezko, [41] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin,
Shinobu Onoda, Hitoshi Sumiya, Vedika Khemani, Curt R. Barends, R. Biswas, S. Boixo, F. G. S. L. Bran-
8

dao, D. A. Buell, et al., “Quantum supremacy using a


programmable superconducting processor,” Nature 574,
505–510 (2019).
[42] S F Edwards and P W Anderson, “Theory of spin
glasses,” Journal of Physics F: Metal Physics 5, 965–974
(1975).
[43] Jonas A. Kjäll, Jens H. Bardarson, and Frank Pollmann,
“Many-body localization in a disordered quantum ising
chain,” Phys. Rev. Lett. 113, 107204 (2014).
[44] Arijeet Pal and David A. Huse, “Many-body localization
phase transition,” Phys. Rev. B 82, 174411 (2010).
[45] D.A. Abanin, J.H. Bardarson, G. De Tomasi,
S. Gopalakrishnan, V. Khemani, S.A. Parameswaran,
F. Pollmann, A.C. Potter, M. Serbyn, and R. Vasseur,
“Distinguishing localization from chaos: Challenges
in finite-size systems,” Annals of Physics 427, 168415
(2021).
[46] B. Foxen, C. Neill, A. Dunsworth, P. Roushan, B. Chiaro,
A. Megrant, J. Kelly, Z. Chen, K. Satzinger, R. Barends,
et al. (Google AI Quantum), “Demonstrating a continu-
ous set of two-qubit gates for near-term quantum algo-
rithms,” Phys. Rev. Lett. 125, 120504 (2020).
[47] Frank Arute, Kunal Arya, Ryan Babbush, Dave Bacon,
Joseph C. Bardin, Rami Barends, Andreas Bengtsson,
Sergio Boixo, Michael Broughton, Bob B. Buckley, et al.,
“Observation of separated dynamics of charge and spin
in the fermi-hubbard model,” (2020), arXiv:2010.07965
[quant-ph].
[48] C. Neill, T. McCourt, X. Mi, Z. Jiang, M. Y. Niu,
W. Mruczkiewicz, I. Aleiner, F. Arute, K. Arya, J. Ata-
laya, et al., “Accurately computing the electronic prop-
erties of a quantum ring,” Nature 594, 508–512 (2021).
[49] This is true up to single-qubit Z rotations on the edge
qubits, if the chain has open boundary conditions. These
could be cancelled by considering the evolution over 4
periods, with minor changes to the result (a cancellation
of terms near the edges). For the sake of simplicity we
will neglect this effect here.
[50] Christian Bartsch and Jochen Gemmer, “Dynamical typ-
icality of quantum expectation values,” Phys. Rev. Lett.
102, 110403 (2009).
[51] Jean Dalibard, Yvan Castin, and Klaus Mølmer, “Wave-
function approach to dissipative processes in quantum
optics,” Phys. Rev. Lett. 68, 580–583 (1992).
9

Supplementary Materials for “Observation of Time-Crystalline


Eigenstate Order on a Quantum Processor”

I. CPHASE GATE IMPLEMENTATION AND a b ≤-5 -4 -3 -2 -1 0


Log10 (P2)
ERROR BENCHMARKING ω10, a 30 Simulation

ω21, a

gmax/2 (MHz)
ε
An essential building block for the quantum circuits 20
used to observe many-body localized DTC is the two-
φ
qubit gate ZZ(φ) = e−i 4 Ẑa Ẑb , which belongs to the more ω10, b
gmax 10
general family of Fermionic Simulation (FSIM) gates hav-
0
ing the unitary form ÛFSIM 41 : g -100 -50 0 50 100
time ε/2 (MHz)

 c 30 d Disorder Instances Average

gmax/2 (MHz)
 6
1 0 0 0 3

 (rad)
4
0 ei(∆+ +∆− ) cos θ −iei(∆+ −∆−,off ) sin θ 0 20

Pauli Error (%)



. 2
ei(∆+ −∆− ) cos θ
 
0 −iei(∆+ +∆−,off ) sin θ 0  10
Simulation 0
2
0 0 0 ei(2∆+ −φ) -50 0 50

gmax/2 (MHz)
6
(3) 30

 (rad)
4
Here θ is the two-qubit iSWAP angle and ∆+ , ∆− and 20
2 1
∆−,off are phases that can be freely adjusted by single- 10 Experiment 0
qubit Z-rotations. In this parametrized representation, -100 -50 0 50 100 0 5 10 15 20
ZZ(φ) = ÛFSIM (θ = 0, ∆− = 0, ∆−,off = 0, φ = 2∆+ ), ε/2 (MHz) Nearest-Neighbor Pair
which is equivalent to a CPHASE gate with conditional-
phase φ and a single-qubit rotation Z( φ2 ) acting on each FIG. 6. Implementing CPHASE gates with tunable
qubit. Precise single-qubit Z-control has already been transmon qubits. a, Schematic of flux pulses used to real-
demonstrated in our previous work33 . Here, we focus on ize a CPHASE gate. The frequencies of two coupled trans-
implementing CPHASE gates with a variable φ. mons, ω10, a and ω10, b , are displaced from their idle positions
The qubits used in our experiment are superconduct- into a configuration wherein ω10, b is detuned from ω21, a by
ing transmon qubits with both tunable frequencies and an amount . At the same time, a flux pulse on the cou-
pler turns on an inter-qubit coupling g > 0 with a maximum
tunable inter-qubit couplings. Due to the existence of
value of gmax for a fixed duration of ∼18 ns. b, Simulated
higher states, a natural way to realize a CPHASE gate values of leakage, P2 , as a function of  and gmax , using typ-
is to bring the |11i and |02i states of two coupled qubits ical device parameters and pulse shapes. c, The values of
close to resonance diabatically, allow the qubits to in- gmax and conditional-phase φ at the locations of minimum
teract for a duration ∼ √ 12 2 , before turning off the leakage, plotted for different values of . Upper panel shows
8g +
simulated results and lower panel shows representative experi-
inter-qubit coupling and ramping the qubits back to their
mental values obtained from one pair of qubits. d, Pauli error
idle configurations. Here |0i and |1i are the ground and rates for each neighboring pair of qubits in the 20-qubit chain
excited states of each qubit, |2i is a leakage state out- used by the experiment, obtained from parallel XEB. Each er-
side the computational space, g denotes the inter-qubit ror rate includes contributions from two random single-qubit
coupling (between the |10i and |01i states) and  is the gates (π/2 rotations around a random axis along the equato-
detuning between the |11i and |02i states during inter- rial plane of the Bloch sphere) and a CPHASE gate. Data are
action. A schematic for the flux pulses to realize the shown for 24 disorder instances, with each instance including
CPHASE gate is shown in Fig. 6a. a different random set of φi across the qubit chain. Red dots
Figure 6b shows simulated values of leakage, namely show the average error of each qubit pair.
the probability of one qubit residing in |2i after the
CPHASE gate, as a function of  and maximum value
of g during interaction, gmax . A narrow arc-like region, is detected for a discrete set of  via |2i state readout
2
corresponding to a contour 8gmax + 2 = constant, can and the corresponding φ is then calibrated using unitary
be seen from the simulation. The values of gmax and φ tomography46 . A polynomial fit is then performed to
along this contour are shown in the upper panel of Fig. 6c, infer values of  and gmax for intermediate values of φ,
where we see a full range of φ ∈ [−2π, 0] can be achieved thereby achieving a continuous family of CPHASE gates
for absolute detuning values of ||/2π < 100 MHz. Since with fully tunable φ. Example experimental calibration
the |01i and |10i states are detuned by ∼200 MHz, their data for , gmax and φ are included in the bottom panel
interaction remains dispersive during the CPHASE gate of Fig. 6c. Excellent agreement with numerical results is
and therefore ensures a small iSWAP angle θ (we confirm found. The discrepancy in values of  likely arises from
this experimentally in the next section). deviation between the actual pulse shapes on the quan-
Experimentally, the leakage-minimizing value of gmax tum processor and those used in the numerical simula-
10

tion. ෡ /X
 = tan−1( Y ෡ )
a b
d cycles d cycles d cycles
To estimate the errors of typical CPHASE gates, we P1,a 1
Qa: |+Xۧ Z () Y Qa: |+Xۧ |0ۧ

FSIM

FSIM

FSIM
perform cross-entropy benchmarking (XEB) similar to 2

previous works33,41 . Here the gates are characterized Qb: |0ۧ Qb: |0ۧ |+Xۧ
d cycles d cycles d cycles
in parallel and therefore include errors arising from any 3 4
Qa: |0ۧ Qa: |+Xۧ |+Xۧ

FSIM

FSIM

FSIM
cross-talk effects. The XEB results for 24 random combi- P1,b
nations of φi across the 20-qubit chain used in the main Qb: |+Xۧ Z () Y Qb: |0ۧ |1ۧ

text are shown in Fig 6d, where we have used the so- 1.0
P1,a
called “cycle” Pauli error as the metric. A cycle Pauli 0.8

Phase ( rad)


P1,b 50
error includes errors from two single-qubit gates and a 0.6 1 +  2
single CPHASE gate. In computing the XEB fidelities,

P1
0 4 − 3
0.4
we have also assumed the CPHASE gate with the cali-
0.2  = -3.495 rad
brated φ as the ideal gate33 . As such, the Pauli errors d=8
− = 0.003 rad
-50
2+ = -3.494 rad
0.0
include contributions from both incoherent effects such
-1.0 -0.5 0.0 0.5 1.0 0 5 10 15 20
as qubit relaxation and dephasing, as well as coherent  ( rad) Number of Floquet cycles, d
effects such as any mis-calibration in φ and residual val- c d
-0.5 -0.37 Control Error:
ues of θ. The error rates are observed to be relatively -2 0.0085 rad
dependent on φ, a likely consequence of changes in co- -0.38

 (rad)

 (rad)
herent and incoherent errors when the pulse parameters -3 -0.39
are varied. Overall, we observe an average error rate of -0.40
-4
0.011, comparable to gates used in our past works33,41 .
-1.5 -0.41
0 5 10 15 20 0 5 10 15 20
Nearest-Neighbor Pair Nearest-Neighbor Pair
II. FLOQUET CALIBRATION OF CPHASE
GATES
FIG. 7. Floquet calibration of single-qubit and con-
ditional phases for CPHASE-like gates. a, Top panel:
After basic pulse-level calibration of CPHASE gates, Gate sequences for calibrating the ∆− angle of the FSIM. Bot-
the ZZ(φ) gate is then calibrated using the technique tom panel: Example experimental data for P1,a and P1,b (|1i
of Floquet calibration47,48 . Floquet calibration utilizes state population for qubits Qa and Qb , respectively) as func-
periodic circuits which selectively amplify different pa- tions of ξ. The circuit depth is fixed at d = 8. Solid black lines
rameters within ÛFSIM , allowing for sensitive detection show fits to a cosine function for each qubit, which allow ∆−
to be extracted. b, Top panel: Gate sequences for calibrating
and rectification of small coherent errors for such quan-
∆+ and φ. For each of the 4 gate sequences, the hX̂i and hŶ i
tum gates. Our past works have primarily focused on projections of the Bloch vector for a given qubit are measured
calibrating the iSWAP-like family of gates, where θ  φ. 
at the end, from which an angle α = tan−1 hŶ i / hX̂i is
For ZZ gates, the opposite limit φ  θ is true and the
optimal calibration circuits are in some cases different computed. Bottom panel: Example experimentally obtained
phase sums (α1 + α2 ) and differences (α4 - α3 ) as functions
from our previous works. In this section, we present the
of d, number of cycles in the Floquet gate sequences. Solid
gate sequences and example calibration results for the black lines show linear fits, the slopes of which determine φ
ZZ gates. For a detailed description of the underlying and ∆+ . c, Experimentally measured φ for each neighboring
theory of Floquet calibration, the reader is directed to pair of qubits in the 20-qubit chain. Results from 40 disor-
our previous publications47,48 . der instances are plotted. The blue and red connected dots
indicate the values of two particular instances, while all other
instances are shown as disconnected purple dots. Dashed lines
A. Calibration of ∆+ , ∆− and φ corresponding to φ = −0.5π and φ = −1.5π. d, Experimental
measurements of φ when a target value is set at −0.4 (dashed
line) for all nearest-neighbor pairs. An average deviation of
The calibration circuits for ∆− are illustrated in 0.0085 rad is found between the target and measured values
Fig. 7a and comprise two Ramsey-like measurements: of φ.
Each qubit is separately initialized along the x-axis of the
Bloch sphere, |Xi. A total number of d FSIM gates are
then applied, which in general rotate the excited qubit −ξb
around the z-axis of the Bloch sphere due to non-zero ting parameters. The value of ∆− is then equal to ξa2d ,
single-qubit phases within the uncalibrated ÛFSIM . At where ξa (ξb ) is the fitted ξ0 for Qa (Qb ). This equiva-
the end of the sequence, a Z gate with a rotation lence may be understood through the fact that 2∆− is
√ angle
ξ is applied to the excited qubit, followed by a Y gate. the difference in the degree of local Z rotations under-
The resulting |1i state population, P1 , is then measured. gone by each qubit after the application of one FSIM gate
Example data for P1 of each qubit are shown in the bot- between them.
tom panel of Fig. 7a, which are fitted to a cosine function The phases ∆+ and φ are calibrated using four periodic
P1 = B0 + B1 cos(ξ + ξ0 ) where B0 , B1 and ξ0 are fit- circuits sharing a similar structure, as indicated in the
11

a d cycles b B. Calibration of θ
Disorder Instances Average
Qa: |1ۧ R (-)

FSIM
P1
Qb: |0ۧ R ( + /2)
0.04 For iSWAP-like gates within the FSIM family, the

iSWAP angle,  (rad)


Data iSWAP angle θ can be accurately calibrated by setting
0.90 Fit
∆− = 0 and applying FSIM gates in succession, which
0.03
0.85 leads to population transfer between the two qubits (ini-
tialized in |10i) with a period of πθ . Discrete fourier trans-
P1

0.80
0.02 form of qubit populations therefore allow θ to be deter-
0.75 mined with very high precision48 . However, CPHASE-
d = 17 like gates typically have small iSWAP angles and such a
0.70  = 0.029 rad 0.01
technique is no longer as effective, since the period for one
-0.4 -0.2 0.0 0.2 0.4 0 5 10 15 20
 ( rad) Nearest-Neighbor Pair
population transfer can be very long and the calibration
data are complicated by noise effects.
FIG. 8. Floquet calibration of small iSWAP angles. a, To circumvent such a problem, we have designed a
Top: Periodic circuit for calibrating θ: Each cycle includes new gate sequence for calibrating small values of θ: Let
an FSIM gate, followed by two single-qubit rotations Rπ (−ψ) us consider the composite unitary ÛCOM = ÛFSIM Xa Yb
and Rπ (ψ + π/2). After d cycles, the excited state population where Xa and Yb are X and Y π-rotations of Qa and Qb ,
P1 of Qb is measured. Bottom: Example experimental data respectively. ÛCOM has the following matrix form:
at a fixed depth d = 17, showing P1 as a function of ψ. Solid
black line shows fit to a sinusoidal function, the amplitude of  
0 0 0 −i
which determines the value of θ. b, Experimentally measured  0 −e−i∆−,off sin θ
θ for each neighboring pair of qubits in the 20-qubit chain. i cos θ 0.
 (4)
Results from 40 disorder instances are plotted as blue dots,
 0 −i cos θ ei∆−,off sin θ 0
and the average value for each qubit pair is plotted as red ie−iφ 0 0 0
dots. Overall, θ has a mean value of 0.022 rad and a standard
deviation 0.014 across all qubit pairs and disorder instances. Here we have set ∆− and ∆+ to 0 for simplicity (non-
zero values of these phases will not affect the calibra-
tion scheme discussed below). With simple algebra, it
can be seen that for qubits initialized in the |10i state,
the excited state population of Qb after applying ÛCOM
top panel of Fig. 7b. For ∆+ , we again separately prepare d times (d being odd) is bounded by two values: For
each qubit in the |Xi state while leaving the other qubit ∆−,off = 0, P1 = cos2 θ ≈ 1 for small values of θ. For
in |0i. The FSIM gate is then applied d times. At the ∆−,off = π2 , P1 = cos2 (dθ). The difference between these
end of the sequence, two tomographic measurements are two values is cos2 θ − cos2 (dθ) which is approximately
performed on the excited qubit to determine  the angle sin2 (dθ) for θ ≈ 0. As such, varying ∆−,off and measuring
−1
of its Bloch vector, α = tan hŶ i / hX̂i . The total the peak-to-peak amplitude of P1 allows determination
accumulated phase α1 + α2 , where α1 (α2 ) is α for Qa of θ. Compared to the fourier-transform technique, the
(Qb ), is equivalent to 2d∆+ . This equivalence arises from method here requires relatively short circuit depth, as a
the fact that 2∆+ physically corresponds to the sum of iSWAP angle of 0.02 rad would yield a peak-to-peak am-
the degrees of local Z rotations on both qubits after the plitude of 0.1 in P1 for d = 17, which can be resolved with
application of one FSIM gate. a relatively small number of single-shot measurements.
The experimental Floquet circuit for calibrating θ is
The conditional-phase φ is calibrated by preparing one shown in the upper panel of Fig. 8a. Here, we have
qubit (Qa ) in |Xi and comparing α when the other qubit injected a variational angle ψ into the single-qubit π-
(Qb ) is initialized in either the |0i or the |1i state. A rotations. Varying ψ effectively changes χ of the FSIM
non-zero φ will cause the two angles, α3 and α4 , to differ gate. The experimental calibration data for a given pair
by an amount α4 − α3 = −dφ46 . Example experimental of qubits are shown in the bottom panel of Fig. 8a, where
values of α1 +α2 and α4 −α3 as a function of d are shown we see oscillations of P1 as a function of ψ. Fitting the
in the bottom panel of Fig. 7b. The slopes of both data data to a sinusoidal function allows a peak-to-peak am-
sets allow ∆+ and φ to be extracted, which are seen to plitude of 0.22 to be determined, which corresponds to a
be very close to the target condition 2∆+ = φ. Figure 7c iSWAP angle of θ = 0.029 rad.
shows experimentally calibrated values of φ across the The iSWAP angles for all qubit pairs of the 20-qubit
20-qubit chain, for a total of 40 disorder instances. It chain are shown in Fig. 8b, where we have included data
can be seen that φ falls within the intended range of for 40 disorder instances in φ. A small average value of
[−0.5π, −1.5π]. Figure 7d shows the calibrated values of 0.022 rad is found for the qubit chain, with the fluctu-
φ when each φi has a target value of −0.4. Comparing ation between disorder instances understood as a result
the measured values of φ with the target value, we find of different detunings between the |01i and |10i states
a small control error of 0.0085 rad for φ. during the flux pulses of different gates.
12

III. DERIVATION OF EFFECTIVE respectively) lie near the edges of the spectrum in all
HAMILTONIANS realizations.
We note that, strictly speaking, a prethermal DTC re-
The bit-string energies shown in the inset to Fig. 3b quires Ĥeff to have a symmetry breaking transition at a
are based on effective Hamiltonians Ĥeff that, in certain finite critical temperature Tc . In this case, ordered initial
limits, approximate the effect of the unitary circuit over states at temperatures T < Tc show long-lived oscilla-
two periods, ÛF2 ≈ e−2iĤeff . Here we derive the relevant tions with an amplitude that depends on the equilibrium
value of the (symmetry breaking) order parameter at
Ĥeff operators for the models in Fig. 3b.
temperature T 29 . As short-ranged models in one dimen-
sion (such as the one under consideration) cannot have
A. Uniform φi = −0.4 order at finite temperature, this model is not prethermal
in the sense we just described. However, thermal correla-
tion lengths may still exceed the size of the system when
For the model with uniform CPHASE angles φi ≡ φ̄ = the latter is small enough. This allows low-temperature
−0.4 and random single-qubit Z rotation angles hi ∈ states of Ĥeff to show long-lived oscillations with a fi-
[−π, π], a period of the time evolution is represented by nite amplitude, even if the equilibrium order parameter
is asymptotically zero for such states.
ÛF0 = Ûz [φ̄, h]Ûx [π − 2] (5)

with
B. Disordered φi ∈ [−1.5π, −0.5π]
−i θ2
P
Ûx [θ] = e n X̂n ,
P In the MBL DTC drive ÛF we set the average
Ûz [φ̄, h] = e−i n (φ̄/4)Ẑn Ẑn+1 +(hn /2)Ẑn . CPHASE angle to φ̄ = −π, which (being ∼ 10 times
larger than in the previous case) breaks the final step in
We have also defined the detuning  = π2 (1 − g); in the
the derivation of Eq. (8). We can however use another
following we take   1, i.e. g close to 1. The evolution
approach, valid if φi = −π + δφi , with |δφi | sufficiently
over two periods is given by
small. In Eq. (7) we replace φ̄ by π +δφ, and noting that
(ÛF0 )2 = Ûz [φ̄, h]Ûx [π − 2]Ûz [φ̄, h]Ûx [π − 2] Ûz [π, 0] = Ûz [−π, 0]49 we have
= Ûz [φ̄, h]Ûx [−2]Ûz [φ̄, −h]Ûx [−2] (6)
P
ÛF2 = Ûz [δφ, 0]ei n (cos(hn )Ŷn −sin(hn )X̂n )(Ẑn−1 +Ẑn+1 )

where we have used the Q comutation properties of the × Ûz [δφ, 0]Ûx [−2] . (10)
perfect π pulse Ûx [π] = n X̂n . Next, we note that
Ûz [φ̄, −h] = Ûz [φ̄, 0]Ûz [0, h]† ; acting by conjugation with If , |δφi |  1, leading-order BCH yields
Ûz [0, h] on Ûx [−2] gives X
Ĥeff = [X̂n + cos(hn )Ŷn (Ẑn−1 + Ẑn+1 )]
P n
2
(ÛF0 )2 = Ûz [φ̄, 0]ei n cos(hn )X̂n +sin(hn )Ŷn
Ûz [φ̄, 0]Ûx [−2] .
δφn 
(7) + Ẑn Ẑn+1 − sin(hn )X̂n (Ẑn−1 + Ẑn+1 ) (11)
4 2
The effective Hamiltonian Ĥeff , satisfying (ÛF0 )2 ≈ The energy of a bit-string state |si is
e−2iĤeff , is then given to leading order in , |φ̄/4|  1 X δφn
via the Baker-Campbell-Hausdorff (BCH) formula: (0)
Es = hs| HF |si = (−1)sn +sn+1 . (12)
n
4
X
Ĥeff = [(1 + cos(hn ))X̂n + sin(hn )Ŷn ]+
n
2 Unlike the result in Eq. (9), this does not single out spe-
cial bit-strings. More specifically, under disorder averag-
φ̄
+ Ẑn Ẑn+1 . (8) ing all bit-strings have the same energy: Es = 0.
4 In our model, the |δφi | angles are not small (they vary
Thus, for any bit-string s ∈ {0, 1}L , the energy of the as- in [−0.5π, 0.5π]) so all orders in BCH should be included
sociated computational basis state |si = |s1 i |s2 i · · · |sL i for an accurate result – the above is only meant to be
is a qualitative approximation. Nevertheless, the indepen-
dence of (disorder-averaged) quantities from the choice
φ̄ X of bit-string can be proven exactly for this model.
Es = hs| Ĥeff |si = (−1)sn +sn+1 , (9) All bit-string states are obtained as |si = X̂s |0i,
4 n
where |0i = |00 . . . 00i is the polarized state and X̂s =
Q
which simply counts the number of “domain walls” (i.e. i:si =1 X̂i flips the qubits where si = 1. We will show
bonds where si 6= si+1 ) in s. Thus the polarized and that all bit-string states give rise to the same dynamics
staggered bit-strings (having 0 and L−1 “domain walls”, as the polarized one, up to a change in the realization of
13

disorder. Indeed, the change of basis that maps |si to |0i (as explained above). Finally we measure in the compu-
acts on ÛF as tational basis and flip the logical value of all bits at even
positions (this is equivalent to acting with P̂π,e at the
X̂s ÛF X̂s = Ûz [φ0 , h0 ]Ûx [πg] (13) final time but avoids any fidelity losses). This way, we
where φ0i= (−1) si +si+1
φi and h0i si
= (−1) hi . φ and 0 manage to effectively invert the circuit without altering
h0 almost define another valid realization of disorder, ex- two-qubit gate fidelities.
cept that wherever si 6= si+1 , we have φ0i ∈ [0.5π, 1.5π]
(as opposed to φi ∈ [−1.5π, −0.5π]). This can be reme-
died by setting φ00i = φ0i − 2π ∈ [−1.5π, −0.5π], and B. Measuring the effect of decoherence
π π π
noting that e−i 2 Ẑi Ẑi+1 ∝ e−i 2 Ẑi e−i 2 Ẑi+1 , so that the
excess 2π CPHASE angle can be transferred to single- Let us model noise as a single-qubit channel
qubit rotations: Ûz [φ0 , h0 ] ∝ Ûz [φ00 , h00 ], where h00i = h0i 
4p

4p
if si−1 = si+1 , or h0i + π otherwise. Thus the dynam- Ep = 1 − I+ D (14)
3 3
ics of bit-string |si under disorder realization (φ, h) is
mapped to the dynamics of |0i under a different real- where I is the identity channel (I(ρ̂) = ρ̂), D is the
ization (φ00 , h00 ). Further, the mapping between realiza- single-qubit erasure channel (D(ρ̂) ∝ I), and p denotes
tions conserves the (uniform) measure over the intervals the error rate. We assume the noise acts symmetri-
φi ∈ [−1.5π, −0.5π], hi ∈ [−π, π]. Thus after averaging cally before and after each iteration of the unitary circuit
over disorder, all bit-strings are equivalent to each other. (while realistically noise acts during the entire evolution,
this is a good approximation as long as p  1). Then the
time evolution of the system over a period is described
IV. ECHO CIRCUIT FOR NOISE MITIGATION by a quantum channel
⊗L ⊗L
The “echo” circuit ÛECHO used to define the normal- ρ̂ 7→ Ep/2 ◦ UF ◦ Ep/2 (ρ̂) ≡ Φ(ρ̂) (15)
ization A0 in Fig. 2c and Fig. 4d of the main text consists
of t steps of forward time evolution under ÛF and t steps where UF (ρ̂) = ÛF ρ̂ÛF† . Similarly the inverted time evo-
for “backward” time evolution under ÛF† . In the absence lution is given by
of noise, the two cancel exactly. Thus deviations from ⊗L
ρ̂ 7→ Ep/2 ◦ UF† ◦ Ep/2
⊗L
(ρ̂) ≡ Φ† (ρ̂) (16)
this outcome quantify the impact of noise.
where UF† (ρ̂) = ÛF† ρ̂ÛF , and the last equality holds be-
A. Circuit inversion cause Ep is self-adjoint. The entire echo circuit sequence
is thus described by the channel (Φ† )t ◦ Φt . The expec-
tation value of Ẑi after the circuit is given by
Our device allows the calibration of a continuous fam-
ily of CPHASE angles on each bond and their use during
 
† t
hẐi iecho
s ≡ Tr Ẑ i (Φ ) ◦ Φ t
(|si hs|) (17)
a single circuit run. Thus it is possible to concatenate
the forward and backward time evolutions ÛFt and (ÛF† )t
The temporal autocorrelator between Ẑi before and after
directly. However, the two-qubit gates in ÛF† would have the echo circuit is simply (−1)si hẐi iecho . Averaging over
s
in general different fidelity than those in ÛF . As a re- all bit-strings yields
sult, the decoherence during ÛECHO would differ from
1 X
that during ÛF . A20 ≡ L (−1)si hẐi iecho
s
To circumvent this, we effectively invert the circuit ÛF 2 s
without changing the two-qubit gates, thus keeping the 1 X h i
fidelity unchanged during the backward time evolution. = L Tr Ẑi (Φ† )t ◦ Φt (Ẑi |si hs|)
2 s
The key idea is to apply X rotations on even qubits,
QL/2
P̂π,e ≡ n=1 X̂2n , before and after each period of the 1 h i
= L Tr (Φt (Ẑi ))2 = kΦt (Ẑi )k2 /kẐi k2 , (18)
circuit that is to be inverted. Indeed conjugating ÛF by 2
P̂π,e changes the sign of all φn CPHASE angles. It also where we have used the definition of adjoint channel,
changes the sign of single-qubit Z rotation angles hn on (V̂ , Φ(Ŵ )) = (Φ† (V̂ ), Ŵ ), relative to the trace inner
even sites. The sign of remaining hn fields and of g, as product (V̂ , Ŵ ) = Tr(V̂ † Ŵ ). Thus from the protocol
well as the relative order of the X and Z parts of the outlined above one extracts the decay of operator norm
drive, can be changed at the level of single-qubit gates, kẐi (t)k ∼ A0 kẐi (0)k which is the leading effect of deco-
with minor effects on fidelity. herence. The ratio A/A0 in Fig. 7d thus gives the overlap
In practice, after t cycles of ÛF , we apply the single- between normalized operators,
qubit rotations P̂π,e only once, and then switch to a uni- !
tary V̂F which has the same 2-qubit gates as ÛF but dif- A Ẑi (0) Ẑi (t)
= . (19)
ferent single-qubit gates chosen so that P̂π,e V̂F P̂π,e = ÛF† A0 kẐi (0)k kẐi (t)k
14

V. SPECTRAL AVERAGES VIA QUANTUM a 1.0 10 1 b


TYPICALITY 10 1

Cumulative fraction of states


10 2
0.8 10 2

/
38,39,50 10 3
Quantum typicality states that, for any observ- 10 3

able Ô in a Hilbert space of dimension 2L , the expecta- 0.6 10 4


0 20 40 60 0 20 40 60
tion value hÔiψ on a random state ψ sampled from the K K
unitarily invariant (Haar) measure obeys these statistical 0.4 K
0
properties: 2
0.2 4
Eψ hÔiψ = hÔi∞ (20) 8
24
1  2  0.0
Varψ hÔiψ = L hÔ i∞ − hÔi2∞ (21) 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.5
2 +1 A A
where hÔi∞ ≡ 2−L Tr(Ô) denotes the expectation value FIG. 9. Simulation of quantum typicality protocol. a,
on the infinite-temperature state. Thus for large L, the Cumulative distribution of autocorrelators A from a set of
matrix element hÔiψ is distributed as a Gaussian cen- 2000 bit-string states pre-processed by a depth-K random
tered at the infinite-temperature value with an extremely circuit ÛS as described in the text, for variable K. We set
narrow standard deviation, ' 2−L/2 , which enables the g = 0.94 (MBL DTC phase). The realization of disorder is
measurement of spectrally-averaged quantities with ex- fixed and A is computed at time t = 30 on qubit Q11 in
ponentially high accuracy from a single pure state. a chain of L = 20 qubits. Inset: relative standard deviation
σ/µ decreases exponentially in K and approaches the random-
state variance (< 2−L/2 ) after depth K ' L. b, Same plot
A. Scrambling circuit and approach to typicality for noisy simulations (depolarizing noise, error rate p = 0.5%
per 2-qubit gate, exact density matrix simulations) of qubit
Q7 in a chain of L = 12 qubits, where we include all 4096
In the main text we report data on spectrally-averaged bit-string states. Inset: relative standard deviation σ/µ. σ
autocorrelators hẐi (0)Ẑi (t)i∞ obtained with a method decreases indefinitely with K due to decoherence, while µ is
based on the idea above, i.e by evaluating hẐi (0)Ẑi (t)iψ not affected.
on a state |ψi which is close to typical random states in
the Hilbert space. In order to prepare a random state √
|+Xi = (|0i + |1i)/ 2 alongside the system qubits Q1 ,
|ψi, we start with a bit-string state and evolve it via a
. . . QL which are initialized in a bit-string state |si. We
scrambling circuit ÛS , as also proposed in Ref.40 . This
evolve the system qubits with the scrambling circuit ÛS
consists of K layers of CZ gates (CPHASE with angle
for depth K, obtaining a joint state |ψisys |+Xia . Then
φ = π) and random single-qubit gates (rotations around a
we apply a CZ gate between the ancilla and system qubit
random direction on the XY plane by an angle θ sampled
i, so that the state “branches” into the superposition
uniformly in [0.4π, 0.6π]). The single-qubit gates vary
randomly in both space and time, so that ÛS is not a 1  
√ |ψisys |0ia + Ẑi |ψisys |1ia (22)
Floquet circuit. After a number of layers K = O(L) (we 2
neglect decoherence for now), the prepared state |ψi =
ÛS |si behaves like typical random vectors in the Hilbert We then evolve the system under the Floquet drive ÛF
space, so that hψ| Ô |ψi = hÔi∞ + δ, where the error δ (a for t periods and again apply a CZ between the ancilla
random variable dependent on the choice of bit-string s and system qubit i, which gives
and of scrambling circuit ÛS ) has zero mean and variance 1  
∼ 2− min(L,cK) for some constant c > 0, i.e., the variance √ ÛFt |ψisys |0ia + Ẑi ÛFt Ẑi |ψisys |1ia (23)
shrinks with increasing K as the state becomes gradually 2
more random, until it saturates to the quantum typicality Finally, we measure the ancilla in the X basis. The ex-
result Eq. (21). In Fig. 9a we show the results of exact pectation value of the measurement is
numerical simulations that confirm this picture. For this
family of random circuits, we find c ' 0.36 (from a fit to 1
hX̂a i = hψ| ÛF−t Ẑi ÛFt Ẑi |ψi + c.c.
the data in the inset to Fig. 9a). 2
= Re hψ| Ẑi (t)Ẑi (0) |ψi
' hẐi (0)Ẑi (t)i∞ (24)
B. Ancilla protocol
where the last line follows from quantum typicality if
To measure two-time correlators hψ| Ẑi (0)Ẑi (t) |ψi in |ψi is random. On a sufficiently large system, and for
the “pseudorandom states” |ψi defined above, we use an large enough K (number of scrambling cycles), this pro-
interferometric protocol similar to the one employed in tocol gives the spectrum-averaged temporal autocorrela-
Ref.33 . We introduce an ancilla qubit initialized in state tor from a single measurement.
15

C. Effect of noise during the scrambling circuit a101 L b L


12 10 1 8
16 10
The above discussion neglects decoherence and treats 20 12
100 10 2
the states during the protocol as pure. Since K must

SG

SG
grow with L for the protocol to succeed, it is especially 15
10 3
0.2
important to understand whether noise during the ran- 10 1 10

SG

SG
5 0.1
dom state preparation process has a negative impact on 10 4
0 0.0
the result. 10 2
0.7 0.8 0.9 1.0 0.7 0.8 0.9 1.0
g g
One can repeat the previous discussion with quantum 10 5
0.7 0.8 0.9 1.0 0.7 0.8 0.9 1.0
channels instead of unitary operators: the system starts g g
in pure state |sihs|sys |+Xih+X|a and evolves under the
scrambling dynamics into ρ̂sys |+Xih+X|a , where ρ̂ = FIG. 10. Numerical simulations of spin glass order pa-
ΦS (|sihs|) and ΦS is a quantum channel representing the rameter. a, Ideal (noiseless) dynamics. χSG averaged over
noisy version of the scrambling circuit ÛS (we neglect even times between t = 50 and t = 60, over initial bit-string
decoherence on the ancilla qubit). The protocol then states and over realizations of disorder. At least 4000 real-
proceeds analogously to the noiseless case and yields the izations are averaged at sizes L = 12 and 16, at least 300 at
final state L = 20. Inset: same data on a linear scale. b, Noisy dy-
namics (exact density matrix simulations). Noise is modeled
1 t by single-qubit depolarizing channels with Pauli error rate
ρ̂0sys,a = Φ (ρ̂)sys |0ih0|a + (Ẑi Φt (Ẑi ρ̂))sys |1ih0|a p = 0.5% after each 2-qubit gate. χSG is averaged over even
2 times between t = 50 and t = 60, over initial bit-string states,
+ (Φt (ρ̂Ẑi )Ẑi )sys |0ih1|a and over realizations of disorder. At least 1000 realizations
are used at sizes L = 8 and 10, at least 100 at L = 12. Inset:
+ (Ẑi Φt (Ẑi ρ̂Ẑi )Ẑi )sys |1ih1|a

(25)
same data on a linear scale.

where Φ is the noisy version of the Floquet evolution ÛF . 0.0 0.2 0.4 0.6 0.8 1.0
ζതr
The expectation of X̂a on this state is 20 1.0
i  [-1.5, -0.5] i = -0.4
0.8
15
1 h
Qubit

i 0.6
hX̂a i = Tr Ẑi Φt (Ẑi ρ̂) + Φt (ρ̂Ẑi )Ẑi

ζതr
10
2 0.4
5 0.2
1
= Tr[{Ẑi (t), Ẑi (0)}ρ̂] (26) 0.0
2 0 20 40 60 80 100 0 20 40 60 80 100 5 10 15 20
t t

where we have defined Ẑi (t) = (Φ† )t [Ẑi ] as the


FIG. 11. Numerical simulations of correlation mea-
Heisenberg-picture evolution of Ẑi with noise. surements. Noiseless simulation of the experiment in Fig. 3d
To see that Eq. (26) approximates the infinite- of the main text. Here we simulate the fractional change in
temperature expectation value hẐi (0)Ẑi (t)i∞ , we observe hẐ(t)i, ζ r (see definition in main text), due to a single bit-
that under a random unitary circuit, noise can be approx- flip at Q11 in initial condition. The simulation is averaged
imated by a global depolarizing channel41 : ΦS (|si hs|) ≈ over 1000 disorder instances for both φi ∈ [−1.5π, −0.5π] and
φi = −0.4.
f K ÛS |si hs| ÛS† + (1 − f K )I/2
ˆ L , i.e. a sum of the ideal
evolution under ÛS and the fully mixed state I/2 ˆ L (Iˆ is
the identity matrix), parametrized by a fidelity f < 1. VI. NUMERICAL RESULTS ON SPIN GLASS
However both the ideal scrambled state ÛS |si and the ORDER PARAMETER
fully-mixed state I/2 ˆ L accurately reproduce the infinite-
temperature expectation value. Thus decoherence dur- Here we show results of numerical simulations of the
ing the random state preparation process may in fact spin glass order parameter used to perform a finite-size
be helpful, rather than harmful (as long as the circuit analysis of the phases in the main text. We define the
ÛS is temporally random, so as to avoid any nontrivial order parameter as
steady states). This is confirmed by the results of ex-
act density matrix simulations of L = 12 qubits in the 1 X0
χSG = hẐi Ẑj i2 (27)
presence of depolarizing noise, in Fig. 9b. The variance L−2
i6=j
between bit-string states falls exponentially in K (depth
of ÛS ) even below the ideal quantum typicality limit of where the primed sum excludes the edges (qubits Q1 and
Eq. (21). The subsequent decay is purely due to deco- QL ) in order to remove the effects of edge modes from
herence: the scrambled state ΦS (ρ̂) asymptotically ap- bulk physics. In a phase with glassy order all the ex-
proaches the fully mixed state I/2 ˆ L as the noisy circuit pectation values hẐi Ẑj i are finite and χSG is extensive
is made deeper. (∼ L). Otherwise, all expectation values asymptotically
16

vanish and χSG /L → 0. anteed to be a quantitatively accurate approximation in


In Fig. 10a we show results of numerical simulations of these structured circuits. Even so, the experimental es-
χSG in the absence of noise, at times t between 50 and timate of the phase transition point 0.84 . gc . 0.88 is
60 cycles, as the length of the qubit chain is scaled from reasonably close to the numerical ones, gc ' 0.83 (with-
L = 12 to L = 20. A finite-size crossing is visible near out noise) and gc ' 0.90 (with noise).
g ' 0.83, separating a side of parameter space (at larger
g) where χSG grows with L, indicative of the MBL-DTC
phase, from one (at smaller g) where χSG decreases with
L, indicative of a thermalizing phase. We also note that VII. NUMERICAL RESULTS OF
the finite-size crossing in these data slowly drifts towards CORRELATION MEASUREMENTS
higher g as t increases (not shown), as expected from slow
thermalization on the ergodic side near the transition.
Repeating the same analysis in the presence of noise The noiseless simulation of the experiment in
yields the data in Fig. 10b. We model noise as a single- Fig. 3d of the main text is shown in Fig. 11.
qubit depolarizing channel with Pauli error rate p = 0.5% Here we generate a total of 1000 disorder instances
acting on both qubits after each 2-qubit gate. Simula- for both φi ∈ [−1.5π, −0.5π] and φi = −0.4.
tions are carried out by exact evolution of the density The values of hẐ(t)i are simulated with the two
matrix, which is memory-intensive and limits the avail- initial conditions |00000000000000000000i (ζ1 ) and
able system sizes to L ≤ 12 within reasonable computa- |00000000001000000000i (ζ2 ), and the ratio ζ r is com-
tional resources. (We use this method rather than quan- puted using the same method as the main text.
tum trajectories51 because the latter method is impracti- It is seen that the ratio ζ r from the noise simulation is
cal for this calculation: as χSG is a nonlinear function of quite similar to the experimentally measured values, de-
the state, each expectation hẐi Ẑj i must be averaged over spite no active error-mitigation for this particular quan-
trajectories separately for each disorder realization). We tity. This is likely attributed to the fact that decoherence
still find a finite-size crossing, though at a considerably introduces a damping factor that is approximately the
higher value of g ' 0.90. We note that the noisy simu- same for both the nominator (|ζ1 − ζ2 |) and denomina-
lations in Fig. 10b do not include the effects of read-out tor (|ζ1 | + |ζ2 |) used to compute ζ r . Consequently, their
error and that the depolarizing noise model is not guar- effects are canceled out after dividing the two quantities.

You might also like