Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Characterizing a non-equilibrium phase transition on a quantum computer

Eli Chertkov,1, ∗ Zihan Cheng,2 Andrew C. Potter,2, 3 Sarang Gopalakrishnan,4 Thomas M.


Gatterman,1 Justin A. Gerber,1 Kevin Gilmore,1 Dan Gresh,1 Alex Hall,1 Aaron Hankin,1 Mitchell
Matheny,1 Tanner Mengle,1 David Hayes,1 Brian Neyenhuis,1 Russell Stutz,1 and Michael Foss-Feig1
1
Quantinuum, 303 South Technology Court, Broomfield, Colorado 80021, USA
2
Department of Physics, University of Texas at Austin, Austin, TX 78712, USA
3
Department of Physics and Astronomy, and Stewart Blusson Quantum Matter Institute,
University of British Columbia, Vancouver, BC, Canada V6T 1Z1
4
Department of Electrical and Computer Engineering,
Princeton University, Princeton, NJ 08544, USA
At transitions between phases of matter, physical systems can exhibit universal behavior inde-
pendent of their microscopic details. Probing such behavior in quantum many-body systems is a
arXiv:2209.12889v1 [quant-ph] 26 Sep 2022

challenging and practically important problem that can be solved by quantum computers, poten-
tially exponentially faster than by classical computers. In this work, we use the Quantinuum H1-1
quantum computer to realize a quantum extension of a simple classical disease spreading process
that is known to exhibit a non-equilibrium phase transition between an active and absorbing state.
Using techniques such as qubit-reuse and error avoidance based on real-time conditional logic (uti-
lized extensively in quantum error correction), we are able to implement large instances of the model
with 73 sites and up to 72 circuit layers, and quantitatively determine the model’s critical proper-
ties. This work demonstrates how quantum computers capable of mid-circuit resets, measurements,
and conditional logic enable the study of difficult problems in quantum many-body physics: the
simulation of open quantum system dynamics and non-equilibrium phase transitions.

A remarkable feature of nature is that complicated sys- energy-conservation does not play a role in many non-
tems governed by completely different microscopic rules, equilibrium processes, the essential features can be cap-
such as water or kitchen magnets, can behave essentially tured by discrete quantum circuits without the large
identically on large length and time scales when chang- overhead for continuous-time Hamiltonian simulation
ing between two phases of matter, as long as they share techniques.
a few basic properties such as symmetry and topology. In this work, we study a dissipative quantum circuit
This universal behavior near phase transitions makes it model that generalizes a canonical classical model, the
possible to predict many behaviors of complex materials contact process2 , known to exhibit a non-equilibrium
by studying simple models. Especially in equilibrium, phase transition in the directed percolation (DP) uni-
physicists have made great progress in understanding versality class3 . We are motivated in part by re-
phase transitions by computing the properties of sim- lated continuous-time models relevant to Rydberg atom
plified models and validating the calculations with ex- quantum simulators, for which a combination of field-
periments. However, thermal equilibrium is an ideal- theoretic4 and numerical5–7 arguments have been put for-
ization that often does not hold; many features of the ward to assess whether quantum fluctuations can result
world around us—from geological formations, to disease in critical properties distinct from classical DP. In large
spreading, to the flocking of birds—arise due to pro- part due to the classical difficulty of obtaining precise
cesses that are fundamentally non-equilibrium in nature. results for sufficiently large system sizes and evolution
Non-equilibrium systems also exhibit universal behav- times, a definitive answer to this question remains elu-
ior at phase transitions, but with a richer and generally sive. We note that even if the universal scaling properties
less-well understood phenomenology than their equilib- of a phase transition in a quantum model turn out to be
rium counterparts. The experimental realization of many classical, verifying such behavior may be classically in-
platforms for simulating quantum dynamics has brought tractable.
a particularly poorly understood question to the fore- We consider a model that can be thought of either
front in recent years: Can microscopic quantum features as a discrete-time version of the continuous-time model
of non-equilibrium systems persist at macroscopic scales studied in Refs.4–7 , or as a quantum generalization of
and impact universal properties of the dynamics? the usual discrete-time classical contact process in which
Quantum computers may be helpful in addressing this branching/coagulation processes are replaced by entan-
question for two reasons. First, classically simulating glement generating controlled rotation gates. Using
open quantum systems can be more difficult than sim- tensor-network methods we attempt to identify and char-
ulating unitary dynamics1 often requiring parametri- acterize the phase transition away from the classical
cally larger memory or simulation time. Second, since point. We then use a trapped-ion quantum computer
to study the model on both sides of the non-equilibrium
phase transition as well as at the approximate critical
∗ eli.chertkov@quantinuum.com point determined from classical numerics. Utilizing the
2

mid-circuit measurement and reset capabilities available tive site at position r by n̂r = |1ih1|r and Et (·) as t
on Quantinuum’s H1 series quantum computer, we are steps of evolution, we examine the active-site density
able to study this model for large times and system sizes hn(r, t)i =Ptr(Et (ρ̂0 )n̂r ), total number of active sites
(72 layers of two-qubit gates acting on a 73-site lattice). hN (t)i = r hn(r, t)i, and mean-squared
P extent of the
The high gate fidelities, together with error avoidance active cluster hR2 (t)i = hN1(t)i r r2 hn(r, t)i.
techniques enabled by real-time (e.g., circuit-level) con- Generically, it is postulated that models with absorb-
ditional logic, allow us to verify power-law growth of ob- ing states and no other explicit symmetries undergo non-
servables near the critical point. Contrary to some find- equilibrium phase transitions in the directed percolation
ings for related continuous-time models5 , our classical (DP) universality class3 . At the critical point p = pc of a
and quantum simulations suggest that classical DP uni- DP transition, observables asymptotically scale as power-
versality is robust to the introduction of quantum fluc- laws: hN (t)i ∼ tΘ and hR2 (t)i ∼ t2/z , where the univer-
tuations. sal exponents Θ and z have been determined from nu-
A driven-dissipative quantum circuit. In this merical studies (Θ = 0.313686(8) and z = 1.580745(10)
work, we study a quantum circuit that we call the one- in 1D)3,10 . Likewise, in 1D, the active-site density pro-
dimensional Floquet quantum contact process (1D FQCP; file at late times and p = pc obeys an asymptotic scaling
see Fig. 1a–c). In each time step, the 1D FQCP circuit form hn(r, t)i ∼ tΘ−1/z f (r/t1/z ), where f (x) is a univer-
executes a layer of probabilistic resets (set each qubit to sal scaling function11 .
|0i with probability p) followed by four alternating layers Implementation on a trapped-ion quantum
i z x,y
of controlled-rotation gates: CRx,y (θ) = e− 4 θ(1+σ̂ )⊗σ̂ computer. We implement the FQCP model on
x y z
where σ̂ , σ̂ , σ̂ are Pauli matrices. Quantinuum’s H1-1 trapped-ion quantum processor12,13
The FQCP is a quantum circuit generalization of the (see Fig. 2d). Using recently developed qubit-
contact process, a simple model for disease spreading reuse techniques involving mid-circuit measurement and
by contact between infected individuals3,8,9 (other quan- resets14–19 , we implement up to t = 18 steps of time
tum generalizations can be found in Refs.4,5 ). In the evolution using 20 qubits, a significant savings over the
1D FQCP, each site of a one-dimensional chain is la- 4t + 1 = 73 qubits needed without qubit-reuse. High-
beled as “active” or “inactive,” corresponding to the fidelity and low-cross-talk mid-circuit resets are particu-
single-qubit states |1i and |0i, respectively. In the cir- larly crucial for this study as the FQCP’s probabilistic re-
cuit, there is a competition between a unitary driv- set channels would have required hundreds of additional
ing and non-unitary dissipative process. The driving ancilla qubits to implement without mid-circuit resets.
(set by θ) caused by the two-qubit gates enables ac- As shown in Fig. 1a, for the single active seed initial
tive sites to spread to inactive neighbors through the state many controlled gates and channels in the model
processes: |10i → cos(θ/2)|10i + eiφ sin(θ/2)|11i where act as identity, and therefore do not need to be applied,
φ ∈ {0, π/2} is a phase (see Fig. 1). For generic choices of reducing the total number of noisy operations executed
θ, φ, such processes thermalize subsystems to the infinite- in experiments. Moreover, for this initial state, the av-
temperature maximally-mixed state ρ̂ ∝ 1̂. We note that eraged dynamics is reflection symmetric about r = 0, so
at θ = π, the model maps bitstrings in the computa- that hn(r, t)i = hn(−r, t)i, which allows us to collect data
tional basis to other bitstrings without generating entan- for all r by measuring only at r ≥ 0, further reducing
glement; at this special point our model is therefore clas- the required number of qubits and gates. Fig. 2a shows
sical, efficiently simulable, and guaranteed to have DP the gates and channels causally connected to qubits mea-
critical exponents (see Fig. 1d and supplement). The sured at r ≥ 0, t > 0 and highlights how only t + 2 physi-
dissipation (set by p) caused by the probabilistic reset cal qubits are needed to execute the model, which would
channels drives the entire system to the product state have required 4t + 1 qubits without exploiting reflection
ρ̂ = |0 . . . 0ih0 . . . 0|, called the absorbing state, which the symmetry (3t + 1 using the symmetry).
system can reach but not exit. In chains of length L, the We utilize an “error avoidance” technique that allows
competition of unitary spreading and spontaneous decay us to significantly reduce the effects of two-qubit errors
separates two regimes: an inactive regime at high de- in our circuits. During each experiment, we maintain a
cay rate p where the active sites decay exponentially in real-time log of qubits known to be currently in the |0i
time, and an active regime where the spreading processes state – from prior mid-circuit resets – and use that log to
produce a finite density of active sites that survives to conditionally avoid the application of any two-qubit gate
long times log t ∼ L. As L → ∞, these regimes become for which the control qubit is known to be in |0i, thereby
sharply distinct phases separated by a critical decay rate avoiding errors accompanying gates that do not induce
pc , at which the active site density grows super-diffusively any dynamics (see Fig. 2b,c).
forming a self-similar fractal active cluster. Results. We experimentally perform dynamics sim-
To probe the transition, we scan p for fixed θ and ulations of the FQCP model for times t = 2, 4, . . . , 18 on
time-evolve an initial state ρ̂0 , that has a single active both sides of the phase transition and near the critical
“seed” at position r = 0 in a chain of −L ≤ r ≤ L point. Notably, from extensive numerical simulations us-
sites, for t time steps and system size L = 2t (see ing matrix product operator (MPO) techniques1,21 (see
Fig. 1a). Denoting the operator that indicates an ac- supplement), we do not find evidence of the non-DP criti-
3

a 4t+1 qubits d

Inactive

b = control-𝑅𝑥 (𝜃) gate


c Branching/Coagulation
𝜃
= control-𝑅𝑦 (𝜃) gate
Decay
p
= Reset to |0⟩ with
probability p ⟿

FIG. 1. A driven-dissipative quantum circuit. a The one-dimensional Floquet quantum contact process evolving an
initial state with a single active site at r = 0 (|1i is active; |0i is inactive). The grayed-out gates and channels act as identity
and do not need to be applied. b The two-qubit gates and quantum channels used in the model. c The two key processes in
the model: quantum branching/coagulation of active sites produced by the controlled-rotation gates and decay of active sites
produced by the probabilistic resets. Active sites (|1i) are shown as yellow circles and inactive sites (|0i) as dark blue circles.
d Sample trajectories at the classical point (θ = π) of the model for p = 0.2 < pc , p = 0.3943 ≈ pc , and p = 0.6 > pc . e The
trajectory-averaged active-site density at θ = 3π/4 for p = 0.1 < pc , p = 0.3 ≈ pc , and p = 0.5 > pc obtained from experiments
executed on the H1-1 quantum computer. Numerical estimates of pc are discussed in supplement.

cal behavior observed in a similar model in Ref.5 . Rather, shown in the heatmaps of Fig. 1e. From these heatmaps,
we find evidence of a directed percolation phase transi- we can qualitatively observe the expected physical behav-
tion for both classical (θ = π, pc ≈ 0.39) and quantum ior: in the active phase (p < pc ) the cluster seeded from
(θ = 3π/4, pc ≈ 0.3) instances of the model. Informed the single active site grows ballistically in size, forming a
by these classical numerics, we perform experiments on clear cone shape; at the critical point (p ≈ pc ) the clus-
the quantum version of the model choosing θ = 3π/4 and ter also grows but sub-ballistically; and in the inactive
targeting (a) the active phase (p = 0.1), (b) the critical phase (p > pc ) the cluster shrinks as the quantum state
regime (p = 0.3), and (c) the inactive phase (p = 0.5). is pulled into the inactive absorbing state.
We employ zero-noise extrapolation22,23 error mitiga-
tion at p = 0.3 (see supplement). We emphasize that, The total number of active sites measured experimen-
while our experiments are not in the quantum advan- tally, compared to the expected theoretical results ob-
tage regime since they involve only 20 qubits that can be tained from noiseless quantum trajectory simulations, are
classically simulated, the dynamics of the FQCP model shown in Fig. 3a. For reference, ballistic (∝ t) and di-
is challenging to compute classically in practice, requir- rected percolation power-law (∝ tΘ ) curves are included.
ing large-scale MPO simulations at times late enough to These plots clearly show how two-qubit gate errors seed
see clear signatures of criticality. new active sites, which leads to a late-time ballistic prop-
agation of errors for p ≈ pc . Despite the gate errors, our
The spreading of active sites as a function of space r data at p = 0.1, 0.5 are quantitatively accurate to late
and time t measured in our three sets of experiments are times. After zero-noise extrapolation, the p = 0.3 data
4

3t+1 qubits t+2 qubits


a 𝜀8
b Conditional gates d
𝜀7 0
𝜀8

𝜀7
𝜀6 𝑧𝑗 𝑧𝑘 𝑞𝑗 𝑞𝑘
𝜀6 Bits Qubits
𝜀5 𝜀5
c Random resets
𝜀4
𝜀4 1

𝜀3
Qubit reuse =
𝜀3

𝑞𝑗
𝜀2 𝜀2 𝑧𝑗 𝑟𝑙 𝑚𝑙 𝑞𝑗
𝜀1 Bits Qubit
𝜀1

FIG. 2. Qubit-reuse and real-time conditional logic. a In our experiments, we measure qubits at positions r ≥ 0 at
the final time, which contain only the depicted gates and channels in their causal cones. By executing the operations in the
highlighted “slices” ε1 , ε2 , . . . and resetting and re-using qubits between slices, we can reduce the number of qubits needed for
the simulation from 3t + 1 to t + 2. For example, in this diagram, we first execute the gates and channels in ε1 , then reset the
qubit at r = −4, and reuse it as the input qubit at r = 2 for ε2 . During the circuit, we use three sets of classical bits zj , rl ,
and ml to log which qubits are known to be |0i from mid-circuit resets, to randomly determine which resets to apply, and to
record mid-circuit measurement results, respectively. b To reduce the effect of two-qubit gate errors, we choose in real-time
not to apply a gate controlled on qubit qj if we know that qj is |0i from prior resets. c We implement each probabilistic
reset channel by applying a reset conditioned on the value of the random bit rl and the logging bit zj . d The H1-1 trapped-
ion quantum computer used in this work holds 20 qubit ions (171 Yb+ , red dots) and 20 sympathetic cooling ions (138 Ba+ ,
white dots). Parallel gating, measurement and reset operations are done in the five zones marked with crossing laser beams.
z z
H1-1 can perform: two-qubit gates (arbitrary-angle e−iφσ̂ ⊗σ̂ /2 gates) with about 3 × 10−3 typical gate infidelity between
any pair of qubits by rearranging ions via transport operations; single-qubit gates with 5 × 10−5 infidelity; state-preparation
and measurement with 3 × 10−3 infidelity; mid-circuit measurements and resets with crosstalk below 1 × 10−4 ; and real-time
conditional logic (see Refs.20 and12 for more details).

is quantitatively accurate up to t ≈ 12, up to which time Discussion. We have demonstrated that near-term
it shows good agreement with DP critical scaling. Simi- quantum computers are capable of studying interesting
larly, Fig. 3b shows the experimentally measured mean- non-equilibrium dissipative quantum phenomena, by ex-
squared extent of the active cluster compared to the the- amining the phase transition of a quantum circuit gen-
oretical results, which also shows a deviation from DP eralization of the contact process. The tools we have
scaling after t ≈ 12 time steps near the critical point. utilized to make this quantum simulation experimentally
Error bars are standard errors of the mean obtained from feasible – such as qubit-reuse and error avoidance using
boot-strap resampling with 100 resamples. real-time conditional logic – are generally applicable and
will be helpful primitives for future studies.
A scaling collapse of the experimental data and the-
oretical simulations at the critical point is shown in Our model, surprisingly, exhibits critical scaling at the
Fig. 4b, with y = hn(r, t)it1/z−Θ plotted versus x = phase transition that differs from recent work on a sim-
r/t1/z and with Θ, z set to the known DP values. For ilar continuous-time model4–6 , according to our tensor-
large t at the critical point, these rescaled points are ex- network numerics. In future work, it would be useful to
pected to lie on a universal curve y = f (x). Even for better understand the origin of this discrepancy. Such a
the limited t available, we are able to see a reasonable study would greatly benefit from the development of ad-
collapse of the data to a universal curve. The largest ditional tools for simulating open quantum systems, since
deviations from the collapse are seen in the t & 12 ex- existing tools1,21,24–31 are currently underdeveloped com-
perimental data, which show clear deviations from the pared to closed system simulation methods.
early-time data, consistent with a change in scaling be- In general, open quantum systems, which can have en-
havior due to TQ gate errors. tropy increasing and decreasing processes, provide a rich
5

a a
p < pc

p ≈ pc

p > pc

b b

FIG. 3. Experimental results from quantum computer. FIG. 4. Scaling collapse on quantum data. a The ex-
a The number of active sites hN (t)i and b the mean-squared perimentally measured active-site density hn(r, t)i versus po-
extent of the active cluster hR2 (t)i versus time on a log-log sition r for fixed times t near the critical point (p = 0.3 ≈ pc ),
scale for p = 0.1, 0.3, 0.5, as measured in experiment (colored mitigated by zero-noise extrapolation. b Scaling collapse of
markers) and obtained from noise-less 10,000-shot quantum the experimental data using known directed percolation ex-
trajectory simulations (black crosses). As a guide to the eye, ponents Θ ≈ 0.31 and z ≈ 1.58. The solid black curve shows
curves showing ballistic (black dotted line) and directed per- the scaling collapsed curve for 10,000-shot noise-less quantum
colation power-law (dashed red line) growth of active sites are trajectory simulations performed at t = 18.
also depicted. The p = 0.3 data has zero-noise extrapolation
(ZNE) applied.
physics or error-robust critical phenomena.

platform for realizing interesting dynamics and steady


state physics and, in practice, are difficult to classically I. ACKNOWLEDGEMENTS
simulate1 . One promising future direction would be to
study open system dynamics in higher spatial dimen- This work was made possible by a large group of
sions on a quantum computer, since classical simula- people, and the authors would like to thank the en-
tions of such dynamics are particularly challenging. It tire Quantinuum team for their many contributions.
would also be important to explore different types of non- The experiments reported in this manuscript were per-
equilibrium phase transitions than the one considered in formed on the Quantinuum system model H1-1 quantum
this work, particularly those that are robust to hardware computer34 , which is powered by Honeywell ion traps.
errors (in our case, the directed percolation transition is Numerical calculations were performed using the ITensor
sensitive to bit flips caused by gate errors). One could library35 . We thank Sebastian Diehl, Michael Buchold,
also investigate the non-equilibrium behavior of quantum Kevin Hemery, Henrik Dreyer, Reza Haghshenas, Natalie
circuits with classical feedback mechanisms32,33 , e.g., Brown, Ciaran Ryan-Anderson, Matthew DeCross, Karl
similar to those employed in quantum error correction, Mayer, Christopher Langlett, Michael Lubasch, Michael
and see if such dynamics can give rise to interesting Wall, Philip Daniel Blocher, Ivan Deutsch, Mohsin Iqbal,
6

and Vedika Khemani for helpful discussions. This re- P. Sloan Foundation through a Sloan Research Fellow-
search was supported in part by the National Science ship. ACP and SG performed this work in part at the
Foundation under Grant No. NSF PHY-1748958. ACP Aspen Center for Physics which is supported by NSF
was supported by DOE DE-SC0022102, and the Alfred grant PHY-1607611.

[1] H. Weimer, A. Kshetrimayum, and R. Orús, Simulation lated spin systems, Phys. Rev. Research 3, 033002 (2021).
methods for open quantum many-body systems, Rev. [16] F. Barratt, J. Dborin, M. Bal, V. Stojevic, F. Pollmann,
Mod. Phys. 93, 015008 (2021). and A. G. Green, Parallel quantum simulation of large
[2] T. E. Harris, Contact interactions on a lattice, Ann. Prob. systems on small NISQ computers, Npj Quantum Inf. 7
2, 969 (1974). (2021).
[3] H. Hinrichsen, Non-equilibrium critical phenomena and [17] E. Chertkov, J. Bohnet, D. Francois, J. Gaebler,
phase transitions into absorbing states, Adv. Phys. 49, D. Gresh, A. Hankin, K. Lee, D. Hayes, B. Neyenhuis,
815 (2000). R. Stutz, A. C. Potter, and M. Foss-Feig, Holographic
[4] M. Marcuzzi, M. Buchhold, S. Diehl, and I. Lesanovsky, dynamics simulations with a trapped-ion quantum com-
Absorbing state phase transition with competing quan- puter, Nat. Phys. 18, 1074 (2022).
tum and classical fluctuations, Phys. Rev. Lett. 116, [18] D. Niu, R. Haghshenas, Y. Zhang, M. Foss-Feig, G. K.-
245701 (2016). L. Chan, and A. C. Potter, Holographic simulation of
[5] F. Carollo, E. Gillman, H. Weimer, and I. Lesanovsky, correlated electrons on a trapped ion quantum processor,
Critical behavior of the quantum contact process in one arXiv:2112.10810 (2021).
dimension, Phys. Rev. Lett. 123, 100604 (2019). [19] Y. Zhang, S. Jahanbani, D. Niu, R. Haghshenas, and
[6] E. Gillman, F. Carollo, and I. Lesanovsky, Numerical A. C. Potter, Qubit-efficient simulation of thermal states
simulation of critical dissipative non-equilibrium quan- with quantum tensor networks (2022), arXiv:2205.06299.
tum systems with an absorbing state, New J. Phys. 21, [20] Quantinuum System Model H1 Product Data Sheet,
093064 (2019). https://www.quantinuum.com/products/h1, Version
[7] M. Jo, J. Lee, K. Choi, and B. Kahng, Absorbing phase 5.00. June 14, 2022.
transition with a continuously varying exponent in a [21] L. Bonnes and A. M. Läuchli, Superoperators vs. trajec-
quantum contact process: A neural network approach, tories for matrix product state simulations of open quan-
Phys. Rev. Research 3, 013238 (2021). tum system: A case study, arXiv:1411.4831 (2014).
[8] J. Marro and R. Dickman, Nonequilibrium Phase Tran- [22] K. Temme, S. Bravyi, and J. M. Gambetta, Error mitiga-
sitions in Lattice Models, Collection Alea-Saclay: Mono- tion for short-depth quantum circuits, Phys. Rev. Lett.
graphs and Texts in Statistical Physics (Cambridge Uni- 119, 180509 (2017).
versity Press, 1999). [23] Y. Li and S. C. Benjamin, Efficient variational quantum
[9] G. Ódor, Universality classes in nonequilibrium lattice simulator incorporating active error minimization, Phys.
systems, Rev. Mod. Phys. 76, 663 (2004). Rev. X 7, 021050 (2017).
[10] I. Jensen, Low-density series expansions for directed per- [24] F. Verstraete, J. J. Garcı́a-Ripoll, and J. I. Cirac, Ma-
colation: I. a new efficient algorithm with applications to trix product density operators: Simulation of finite-
the square lattice, J. Phys. A 32, 5233 (1999). temperature and dissipative systems, Phys. Rev. Lett.
[11] H. Hinrichsen, Non-equilibrium phase transitions, Phys- 93, 207204 (2004).
ica A 369, 1 (2006). [25] J. Cui, J. I. Cirac, and M. C. Bañuls, Variational ma-
[12] C. Ryan-Anderson, N. C. Brown, M. S. Allman, B. Arkin, trix product operators for the steady state of dissipative
G. Asa-Attuah, C. Baldwin, J. Berg, J. G. Bohnet, quantum systems, Phys. Rev. Lett. 114, 220601 (2015).
S. Braxton, N. Burdick, J. P. Campora, A. Chernoguzov, [26] E. Mascarenhas, H. Flayac, and V. Savona, Matrix-
J. Esposito, B. Evans, D. Francois, J. P. Gaebler, T. M. product-operator approach to the nonequilibrium steady
Gatterman, J. Gerber, K. Gilmore, D. Gresh, A. Hall, state of driven-dissipative quantum arrays, Phys. Rev. A
A. Hankin, J. Hostetter, D. Lucchetti, K. Mayer, J. My- 92, 022116 (2015).
ers, B. Neyenhuis, J. Santiago, J. Sedlacek, T. Skripka, [27] A. H. Werner, D. Jaschke, P. Silvi, M. Kliesch,
A. Slattery, R. P. Stutz, J. Tait, R. Tobey, G. Vittorini, T. Calarco, J. Eisert, and S. Montangero, Positive tensor
J. Walker, and D. Hayes, Implementing fault-tolerant en- network approach for simulating open quantum many-
tangling gates on the five-qubit code and the color code, body systems, Phys. Rev. Lett. 116, 237201 (2016).
arXiv:2208.01863 (2022). [28] C. D. White, M. Zaletel, R. S. K. Mong, and G. Refael,
[13] J. M. Pino, J. M. Dreiling, C. Figgatt, J. P. Gaebler, Quantum dynamics of thermalizing systems, Phys. Rev.
S. A. Moses, M. S. Allman, C. H. Baldwin, M. Foss-Feig, B 97, 035127 (2018).
D. Hayes, K. Mayer, C. Ryan-Anderson, and B. Neyen- [29] D. Jaschke, S. Montangero, and L. D. Carr, One-
huis, Demonstration of the trapped-ion quantum CCD dimensional many-body entangled open quantum sys-
computer architecture, Nature 592, 209 (2021). tems with tensor network methods, Quantum Sci. Tech-
[14] I. H. Kim, Holographic quantum simulation, nol. 4, 013001 (2018).
arXiv:1702.02093 (2017), arXiv:1702.02093. [30] S. Cheng, C. Cao, C. Zhang, Y. Liu, S.-Y. Hou, P. Xu,
[15] M. Foss-Feig, D. Hayes, J. M. Dreiling, C. Figgatt, J. P. and B. Zeng, Simulating noisy quantum circuits with ma-
Gaebler, S. A. Moses, J. M. Pino, and A. C. Potter, trix product density operators, Phys. Rev. Research 3,
Holographic quantum algorithms for simulating corre- 023005 (2021).
7

[31] Z. Cheng and A. C. Potter, A matrix product opera- [33] T. Iadecola, S. Ganeshan, J. H. Pixley, and J. H. Wilson,
tor approach to non-equilibrium floquet steady states, Dynamical entanglement transition in the probabilistic
arXiv:2206.07740 (2022). control of chaos, arXiv:2207.12415 (2022).
[32] M. Buchhold, T. Müller, and S. Diehl, Revealing [34] Quantinuum H1-1, https://www.quantinuum.com/, May
measurement-induced phase transitions by pre-selection, 25-August 22, 2022.
arXiv:2208.10506 (2022). [35] M. Fishman, S. R. White, and E. M. Stoudenmire, The
ITensor Software Library for Tensor Network Calcula-
tions, SciPost Phys. Codebases , 4 (2022).
Supplementary Information:
Characterizing a non-equilibrium phase transition on a quantum computer
Eli Chertkov,1, ∗ Zihan Cheng,2 Andrew C. Potter,2, 3 Sarang Gopalakrishnan,4 Thomas M.
Gatterman,1 Justin A. Gerber,1 Kevin Gilmore,1 Dan Gresh,1 Alex Hall,1 Aaron Hankin,1 Mitchell
Matheny,1 Tanner Mengle,1 David Hayes,1 Brian Neyenhuis,1 Russell Stutz,1 and Michael Foss-Feig1
1
Quantinuum, 303 South Technology Court, Broomfield, Colorado 80021, USA
2
Department of Physics, University of Texas at Austin, Austin, TX 78712, USA
3
Department of Physics and Astronomy, and Stewart Blusson Quantum Matter Institute,
University of British Columbia, Vancouver, BC, Canada V6T 1Z1
4
Department of Electrical and Computer Engineering,
Princeton University, Princeton, NJ 08544, USA
arXiv:2209.12889v1 [quant-ph] 26 Sep 2022

CONTENTS I. PROPERTIES OF THE MODEL

I. Properties of the model 1 In this work, we numerically and experimentally probe


A. Space-time symmetries 1 the one-dimensional Floquet quantum contact process
B. Internal symmetries 1 (FQCP), a quantum circuit containing elements that
drive the system to infinite temperature and dissipative
C. Connections to other ergodic phenomena 2
elements that push the system into an absorbing state
∣0⋯0⟩. Below we detail some of the properties of this
II. Methods 2 model.
A. Implementation details 2
B. Error mitigation 3
C. Experimental resource requirements 3 A. Space-time symmetries
D. Effects of hardware errors 4
The FQCP is periodic in time and executes a full Flo-
III. Additional experimental data 5 quet period in t = 2 time steps. Within a time step,
controlled rotation gates are applied left and right with
IV. Classically solvable point numerical simulations 6 control qubits on the odd sublattice followed by left and
1. Classical nature of θ = π point 6 right with control qubits on the even sublattice, resulting
2. Details of numerics 7 in four layers of two-qubit gates between times t and t+1.
3. Simulation results and determination of Note that two identical controlled-unitary gates with the
critical point properties 7 same control qubit or the same target qubit commute
with one another. This implies that commensurate neigh-
boring layers of left- and right-facing controlled-rotation
V. Quantum point numerical simulations 7
gates can be commuted past one another (e.g., the first
A. Dynamics 7 and second layers or third and fourth layers between t = 0
1. Details of numerics 7 and t = 1 in Fig. 1a). Therefore, the model is symmet-
2. Analysis of simulation errors 8 ric under inversion r → −r (for open boundary condi-
3. Simulation results and determination of tion chains of odd length or periodic boundary condition
critical point properties 10 chains of even length). While we only study the model
4. Validation of methods on classical point 10 with open boundary conditions, the model with periodic
5. Dynamics of entanglement entropies and boundary conditions is spatially periodic with a two-site
other quantities 13 unit cell.
In this work, we have focused on the FQCP applied to
an initial state ∣0⋯010⋯0⟩ with a single active site at the
B. Open DMRG simulation of quasi-steady
center site r = 0 of the chain. This state breaks transla-
state 13
1. Details of the method 13
tional invariance, but does preserve inversion symmetry.
2. Critical point scaling of relaxation time 15

References 17 B. Internal symmetries

In additional to temporal and spatial symmetries,


quantum many-body systems can have internal symme-
∗ eli.chertkov@quantinuum.com tries, such as charge conservation. When such symme-
2


/HYHOVSDFLQJUDWLR r level-spacing ratios were computed using full exact diag-
onalization on systems with open boundary conditions
and an even number of sites (so these models do not pos-
 )4&3 XDQGYURWDWLRQV sess inversion symmetry, which requires an odd number
)4&3 DOOXURWDWLRQV of sites).
3RLVVRQ
 &2(

C. Connections to other ergodic phenomena



      As we showed in the previous section, in the p = 0
6\VWHPVL]HL dissipation-less limit, the FQCP is an ergodic (thermal-
izing) unitary Floquet circuit. This means that generic
FIG. S1. Level-spacing statistics. The average nearest- eigenstates of the Floquet unitary appear as volume-law
neighbor level-spacing ratio r = min(Ej+1 − Ej , Ej+2 − infinite-temperature eigenstates. However, by construc-
Ej+1 )/ max(Ej+1 − Ej , Ej+2 − Ej+1 ) of the Floquet unitary tion, the completely unentangled absorbing state ∣0 . . . 0⟩
U with eigenvalues e−iEj (sorted so that Ej ≤ Ej+1 ) for the is also an eigenstate of the Floquet unitary. Therefore,
FQCP with no resets (p = 0) for different system sizes L. the absorbing state is a Floquet quantum many-body
The approximate value expected for Poisson (COE) statistics scar10,11 , an athermal eigenstate of the thermalizing Flo-
is marked with a dashed black line (dotted red line). The quet unitary that does not possess extensive entangle-
model as described in the main text, which contains X and
ment entropy.
Y rotations, has level-spacing statistics consistent with COE,
while the model with only X rotation appears consistent with
The FQCP is also closely related to recently stud-
Poisson level-spacings. ied measurement-induced phase transition (MIPT)
models12–14 , quantum circuits with unitary gates inter-
spersed with random measurements that demonstrate a
tries exist, they can obstruct the ability of the system phase transition in the entanglement entropies of quan-
to thermalize according to the eigenstate thermalization tum trajectories. The FQCP can be thought of as essen-
hypothesis1–3 . Here we provide numerical evidence that tially an MIPT model with a reset following each mea-
the FQCP model does produce thermalizing behavior and surement and with a specially constrained set of unitary
so does not contain hidden internal symmetries. gates (controlled rotation gates instead of Haar random
The FQCP contains both X and Y axis controlled ro- or Clifford gates). For this reason, it is likely that the
tations. This differs from the structure of the (Hamil- FQCP undergoes an MIPT transition. It would be in-
tonian) quantum contact process studied in previous teresting future work to validate whether such an MIPT
work4–7 , which only contained X axis rotations. We transition exists in the model and whether it coincides ex-
chose this particular structure to ensure that the model, actly with the directed percolation transition or whether
at p = 0 when it is a non-dissipative unitary model, is it occurs deeper in the active phase.
not accidentally at a fine-tuned integrable point, which
could cause it to display non-generic (and potentially
non-thermalizing) behavior. II. METHODS
One diagnostic used to assess whether a Hamiltonian
(or Floquet unitary) model is integrable or non-integrable Here we present additional details on the methods uti-
is the average level-spacing ratio ⟨r⟩ between the spacings lized in our FQCP experiments performed on the H1-1
of neighboring energy levels (or Floquet eigenvalues)8 . quantum computer.
Models that possess many conserved quantities, such
as integrable models, have eigenvalues that do not re-
pel one another, which causes them to have an average A. Implementation details
level-spacing ratio of ⟨r⟩P oisson ≈ 0.39. On the other
hand, ergodic models without symmetries generally have Each controlled-rotation gate in the FQCP model
average level-spacing ratios corresponding to values ob- is realized using a single arbitrary-angle Rzz (φ) =
tained from random matrix theory, such as ⟨r⟩COE ≈ 0.53 exp(−iφσ̂ z ⊗ σ̂ z /2) gate that is native to the H1-1 de-
for the circular orthogonal ensemble (COE) of random vice, as shown in Fig. S2a. These gates, as well as the
matrices8,9 . random resets in the model, are conditionally applied
As shown in Fig. S1, the all-X-rotation version of the based on the values of classical bits zj , rl (see Fig. 2b,c).
FQCP (at p = 0) appears to possess Poisson level-spacing The zj bits are used to track in real-time which of the
statistics, while the X and Y rotation model is consis- qubits are known to be in the ∣0⟩ state due to prior mid-
tent with COE level-spacing statistics. This suggests circuit resets. Before each conditional reset, we perform
that the all-X-rotation model has unknown conserved a conditional measurement, which we record to bit ml .
quantities and that the X-and-Y -rotation FQCP model After each conditional gate and reset, we update the zj
has no conserved quantities and is ergodic. The average bits to track the locations of the ∣0⟩ states. The rl bits
3

a Native TQ gate b RNG c Zero-noise extrapolation


𝐻
𝜙
𝑋
𝜃 𝑅𝑧𝑧 −𝜃/2
= 𝜙 𝜙
𝑅𝑧 𝜃/2 𝐻 𝑅𝑥 𝜃𝑝
𝑋
𝑞1 𝑞2 𝑞1 𝑞2 𝑞1 𝑞2
𝑞 𝑟𝑙 𝜙
𝑞1 𝑞2

FIG. S2. Implementations of gates, random number generation, and zero-noise extrapolation. a The gate decom-
position of a controlled-Rx (θ) gate in terms of the native Rzz (φ) two-qubit gate available in the H1-1 device. b At the beginning

of each quantum circuit, we generate random bits rl that are 1 with probability p using the quantum computer, by performing
single qubit Rx (θp ) = exp(−iθp σ̂ x /2) rotations followed by measurements, where θp = 2 arcsin( p). We can generate arbitrarily
many random bits this way by resetting the qubit and repeating the procedure. c When performing zero-noise extrapolation,
we replace each Rzz (φ) gate with three Rzz (φ) gates and two single-qubit σ̂ x gates.

are random bits used to trigger the application of the circuit for t = 2, 4, . . . , 18 without zero-noise extrapola-
mid-circuit resets. As shown in Fig. S2b, the rl bits are tion.
generated by the quantum computer itself at the begin-
ning of each circuit, using single-qubit gates, mid-circuit
measurements, and mid-circuit resets. C. Experimental resource requirements
To save run-time, we do not apply the first layer of
random resets at t = 0. This is because with probability As shown in Fig. 2 and discussed in the main text, the
p the random reset at r = 0, t = 0 is applied to the ∣1⟩ state FQCP circuits executed on the H1-1 quantum computer
at r = 0, causing the dynamics to immediately fall into utilized qubit-reuse as well as conditional quantum gates,
the absorbing state. After this time step, we know that measurements, and resets that are applied depending on
the state remains unchanged for the rest of the circuit. To the values of classical bits that are updated in real-time
avoid simulating this trivial dynamics on the device, we during the calculation. As such, the number of times
do not apply the first random reset and simply reweight each conditional operation is triggered varies per circuit.
measured observables by 1 − p, the probability of there Here we describe the number of resources – qubits, gates,
not being an r = 0 reset at t = 0. At the final time t > 0 etc. – utilized on average in these circuits.
in each experiment, we measure all r ≥ 0 qubits; these The number of qubits used in our calculations varied
measurements as well as all mid-circuit measurements between 9 and 20, the maximum number of qubits avail-
are performed in the σ̂ z -basis. able on H1-1 at the time of these experiments17 . When
performing qubit-reused calculations of the form shown
in Fig. 2, one is free to choose how many of the high-
B. Error mitigation lighted sets of operations (ε1 , ε2 , . . .) to perform in par-
allel before resetting and re-using qubits. For example,
To reduce the effects of errors on our measured re- one could execute the gates and channels in ε1 , ε2 , then
sults, we implement an error mitigation scheme known reset the qubits at r = −2t, −2t + 1, then re-use them as
as zero-noise extrapolation15,16 . We apply zero-noise ex- input qubits at r = 2, 3. Doing more than one set in
trapolation by performing two sets of experiments, one parallel increases the number of qubits required for the
for the original model (TQ error q) and one for a mod- calculation, but speeds up the calculation by increasing
ified model with each TQ gate replaced by three TQ the number of TQ gates that can be executed in paral-
gates (effective TQ error 3q; see Fig. S2c). Using these lel. We choose to perform the qubit-reuse in such a way
two data sets, we extrapolate observables to the zero- so that no qubits are reset and reused until at least 10
noise limit by performing a simple linear extrapolation qubits participate in the calculation. We do it this way
O(q = 0) = O(q) − 12 (O(3q) − O(q)) = 3O(q)/2 − O(3q)/2. because the H1-1 quantum computer can perform up to 5
In our experiments at p = 0.3, for t = 2, 4, 6, . . . , 14 we parallel TQ gates at once with its 5 gate zones17 . Given
perform zero-noise extrapolation and collect 600 shots of this qubit-reuse strategy, the required number of qubits
data for the original circuit as well as 200 shots for the for the circuits simulating t steps of time evolution is as
magnified-error circuit (except for the t = 12, 14 circuits shown in Fig. S3a.
which had 190, 100 shots for the magnified-error circuit, For each circuit executed on the quantum computer,
respectively). For the p = 0.1, 0.5 experiments, we per- there is a different pattern of random resets and therefore
form 100 shots of the circuit for t = 2, 4, . . . , 18 and do a different number of gates, measurements, and resets
not use zero-noise extrapolation. For the p = 0.3 density activated by the real-time conditional logic. Fig. S3b–
profile shown in Fig. 1e, we performed 100 shots of the e shows the statistics of the conditional operations.
4

b c

d e

FIG. S3. Resources used in FQCP circuits. a The number of qubits used for FQCP circuits with time evolution up to t. b
The average number of activated two-qubit (TQ) gates versus t for 100 shots of p = 0.3 circuits (the total number of conditional
TQ gates is marked with a red dashed line). c The fraction of TQ gates that are activated. d The average number of activated
mid-circuit measures/resets, excluding measures/resets used in random number generation (the total number of conditional
measures/resets is marked with a red dashed line). e The fraction of mid-circuit measures/resets that are activated.

Roughly, as a first-order approximation, we expect that stroy the absorbing state, and thereby the signatures of
≈ p of the two-qubit (TQ) controlled-rotation gates are the DP transition, at late times.
not activated due to their control qubit being reset in the The largest error source in the H1-1 quantum com-
time step before the gate is applied. This is consistent puter, and one that directly affects the absorbing state,
with our findings in Fig. S3c, though in reality the frac-
tion of activated TQ gates is less than 1 − p because of
is the two-qubit (TQ) gate, which has an average gate
infidelity of 3 × 10−3 (see Ref.17 for a recent description
higher-order processes in which many random resets con- of the errors in the device). Ideally, the TQ controlled-
spire to prevent the execution of TQ gates. Naively, we rotation gates in the FQCP do not create (or annihilate)
would expect that exactly p of the of the conditional mid- active sites if the control qubit is in the ∣0⟩ state. In par-
circuit measures/resets should be applied since these con- ticular, each TQ gate should leave the absorbing state
ditional operations are activated by classical bits specif- unaffected. However, TQ gate errors break this fine-
ically sampled from a random number generator to be 1 tuned property, leading to the creation of active sites,
with probability p. However, it is possible for mid-circuit even if the absorbing state has been reached. We uti-
measures/resets to not be activated if it is known that
the qubit is already in the ∣0⟩ state from prior resets (see
lize real-time conditional logic to only apply TQ gates as
necessary to avoid the damaging effects of unnecessarily
Fig. 3c), which makes the actual activated fraction less applied TQ gates (see Fig. 2). This “error avoidance”
than p. Fig. S4 shows example realizations of the condi- technique leads to a ≈ 30% reduction in the number of
tional operations for different reset patterns. TQ gates used near the critical point (see Fig. S3b).
Another source of error in H1-1 is a dephasing er-
ror that accumulates on all qubits during idling and ion
D. Effects of hardware errors transport. While such dephasing errors will be detri-
mental to quantum coherences, they will not necessarily
The ideal FQCP model has a perfect absorbing state destroy the absorbing state property like the TQ gate er-
∣0⋯0⟩, which cannot be escaped after it is reached. This rors. Since these dephasing errors are essentially z-basis
property of the model ensures the directed percolation rotations, they do not directly affect the ∣0⋯0⟩ absorb-
(DP) universality of the phase transition. However, in ing state. They can, however, create active sites that
our experiments, the model is subject to errors that de- destroy the absorbing state if the dephasing occurs dur-
5

ing the execution of a TQ controlled-rotation gate. As


shown in Fig. S2a, the controlled-rotation gates are im-
a plemented with a Rzz gate surrounded by single-qubit
gates. If there is a significant delay between the appli-
cation of the single-qubit gates before and after the Rzz
gate, then a significant dephasing error can accumulate
within the conditional rotation gate, leading to an effec-
tive TQ gate error that creates active sites. To mitigate
this effect, we submit our circuits to the quantum device
in such a way so that this within-TQ-gate delay time is
minimized.
In principle, others errors such as reset, measurement,
reset crosstalk, and measurement crosstalk errors can
also lead to the creation of active sites, though in the
H1-1 device these effects are significantly less important
than the effects of TQ gate and dephasing errors.
b
III. ADDITIONAL EXPERIMENTAL DATA

From our experimental data, we were able to measure


observables not discussed in the main text. These include

P (t) ≡ 1 − tr(Et (ρ̂0 )∣0⋯0⟩⟨0⋯0∣)


= 1 − ⟨0⋯0∣Et (ρ̂0 )∣0⋯0⟩,
Pr≥0 (t) ≡ 1 − trr<0 (Et (ρ̂0 )∣0r=0 ⋯0r=2t ⟩⟨0r=0 ⋯0r=2t ∣),
(S1)

which are the survival probability and right-side-survival


probability, respectively. These observables directly
probe the probability of a cluster seeded by a single ini-
c tial state surviving for a time t. At the critical points of
directed percolation (DP) phase transitions, the survival
probability has been observed to scale asymptotically as
P (t) ∼ t−δ , where δ ≈ 0.159464 in one dimension18,19 .
We suspect that Pr≥0 (t) should also obey the same scal-
ing at the critical point. Due to reflection symmetry
r ↔ −r, if Pr≥0 (t) = 0 then P (t) = 0; however, in general
Pr≥0 (t) ≤ P (t).
These survival probabilities are computed from the
same set of σ̂ z -basis measurements used to gather the ex-
perimental data presented in the main text. Since these
quantities involve measuring the states of all r or r ≥ 0
qubits at a fixed time t, most of the mid-circuit mea-
surement data from the experiments cannot be used to
compute P (t) or Pr≥0 (t); only final-time measurement
FIG. S4. Examples of conditionally executed quantum
data in which many qubits are measured can be used to
operations. Examples of the quantum operations executed
compute them. In our circuits (depicted in Fig. 2a), we
measure all −2t ≤ r ≤ 2t qubits for times t ≤ 9 and mea-
for different experimentally realized random reset patterns
for t = 4 steps of time evolution at a p = 0.1, b p = 0.3, c
p = 0.5. Conditional operations that are activated are colored sure 0 ≤ r ≤ 2t qubits for time t ≤ 18. Therefore, we
black and orange, that are not activated are colored light blue, collected P (t) data up to t = 9 and Pr≥0 (t) data up to
and that are not executed (even conditionally) are colored t = 18.
light gray. The random reset pattern is indicated by the ∣0⟩ Fig. S5a,b show the experimentally measured survival
text. The numbers next to the boxes indicate the recorded and right-side-survival probabilities, respectively, com-
mid-circuit measurement results obtained in experiment for pared to the theoretical expectation from ideal simu-
the given reset pattern. Note that even if a measure is not lations (with zero-noise extrapolation utilized for p =
performed in hardware, a 0 result can be recorded if a qubit
is known to be in the ∣0⟩ state due to a prior reset.
0.3). In Fig. S5c, we also plot the center-site density
⟨n(r = 0, t)⟩ which at a 1D DP transition is expected
6


([SHULPHQWp = 0.3 [74JDWHV 7KHRU\
a  ([SHULPHQWp = 0.3 [74JDWHV t
t
([SHULPHQWp = 0.3 =1(



N(t)




    
t
b FIG. S6. Zero-noise extrapolation of observable. The
experimentally measured total number of active sites ⟨N (t)⟩
for FQCP circuits with the original number (“1x”) and three-
times-as-many (“3x”) two-qubit (TQ) gates, as well as the
zero-noise extrapolated (ZNE) values. For comparison, bal-
listic (black dotted lines) and DP power-law (red dashed lines)
scalings are shown, as well as noise-free ideal simulations
(“Theory”). Showing data for p = 0.3 and θ = 3π/4.

to scale as ⟨n(r = 0, t)⟩ ∼ tΘ−1/z for Θ = 0.313686 and


z = 1.58074518,19 . Though the data are limited to short
times and the error bars are large due to limited sam-
pling, the p = 0.3 data for all of the measured observables
are mostly consistent with the expected DP power-law
scalings. Notably, the Pr≥0 (t) data for t ≳ 12 appears
to deviate from the expected behavior due to TQ gate
c errors.
Fig. S6 shows the effect of zero-noise extrapolation per-
formed for the ⟨N (t)⟩ observable. For other observables,
any improvements in accuracy due to ZNE were not re-
solvable within the error bars due to limited sampling.

IV. CLASSICALLY SOLVABLE POINT


NUMERICAL SIMULATIONS

As discussed in the main text, the dynamics of the


FQCP at θ = π can be simulated efficiently classically by
evolving a Markov chain. Here we describe our numerical
simulations at this “classically solvable” point.

FIG. S5. Survival probabilities and center-site density. 1. Classical nature of θ = π point
The experimentally measured a survival probability P (t), b
right-side-survival probability Pr≥0 (t), and c center-site den-
sity ⟨n(r = 0, t)⟩ for the Floquet quantum contact process When θ = π, the controlled x- and y-rotation gates be-
with p = 0.1, 0.3, 0.5 and θ = 3π/4. The p = 0.3 data includes come −i times controlled-X (CX) and controlled-Y (CY)
zero-noise extrapolation (ZNE), as described in the main text. gates, respectivly. When a bitstring is provided as in-
put to one of these gates, a bitstring (times a phase
±1, −i) is returned as output. The probabilistic reset
channel also has this property of mapping bitstrings to
bitstrings when considering the behavior of each z-basis-
7

measurement quantum trajectory. Therefore, we see that times when p deviates substantially from this value. One
when θ = π and the initial state is a bitstring, each quan- procedure for estimating the critical point pc is to esti-
tum trajectory stays as an unentangled bitstring product mate where the late-time effective exponents as a func-
state during its entire time evolution (up to an irrelevant tion of p cross. That is, where δO (t; p) = δO (t′ ; p) for
phase). t ≠ t′ , t, t′ ≫ 1. Fig. S8 shows the effective exponents
of ⟨N (t)⟩ versus p for fixed t, showing that they cross
near p = 0.3944 (the inset shows a zoomed-out view). At
2. Details of numerics this crossing point, the effective exponent is 0.21% away
from the known DP value. Note that the crossings of
The dynamics of bitstrings due to applications of the effective exponent curves drift with t. For the classi-
CX/CY gates and random resets can be simulated highly cal point, we are able to work at large enough t so that
efficiently classically. We simulate the evolution starting this drift is barely detectable. However for the quantum
from a single seed initial state 0⋯010⋯0 and work in an point, where we perform a similar analysis (see below),
effectively infinite system-size limit. We do this by stor- we are only able to perform simulations at small t and so
ing in memory the locations of the 1’s in the current state need to carefully examine the finite-t dependence of the
(in a set, or hash table) and updating these locations af- effective exponent crossings.
ter each CX/CY gate. When applying a random reset
channel at site j, we compute a random number r from
a pseudo-random number generator and remove the 1 at V. QUANTUM POINT NUMERICAL
site j from the set if r < p. This process amounts to SIMULATIONS
sampling a quantum trajectory from the channel. Be-
fore each layer of reset channels, we measure observables For θ ≠ π, quantum superposition and entanglement
based on the current state and keep running averages of affect the physics of the FQCP. In this work, we focus
the observables and their variances. Our implementation on a single point θ = 3π/4, which we call the quantum
of this simulation is written in Julia (version 1.6.2)20 . point. At the quantum point, we use matrix-product
operator (MPO) based methods to simulate the dynamics
and determine the quasi-steady state of the model. From
3. Simulation results and determination of critical point these simulations, we determine the approximate critical
properties point and critical exponents in the model, which inform
the choice of parameters used in our experiments.
We perform t = 1, . . . , 1000 time step simulations from
the initial seed state for many values of p ∈ [0.35, 0.45],
repeating each simulation up to 1.12×107 times to obtain A. Dynamics
small error bars on numerical quantities. We parallelize
our calculations over many nodes of a high-performance 1. Details of numerics
computing cluster. From these simulations we compute
the observables ⟨N (t)⟩, ⟨R2 (t)⟩, P (t), ⟨n(r = 0, t)⟩, since In our dynamics simulations, the time-evolved state
they are all diagonal in the σ̂ z bitstring basis. is a mixed state density matrix represented as a matrix
For each observable O(t), we also compute an “effec- product operator
tive exponent”
[1] [2] [L]
d−1
log[O(t + dt)/O(t)] log[O(t + dt)] − log[O(t)] ρ= ∑ As1 ,s′ As2 ,s′ ⋯AsL ,s′ ∣s1 ⟩⟨s′1 ∣⊗⋯⊗∣sL ⟩⟨s′L ∣
δO (t) = = s1 ,s′1 ,...,sL ,s′L =0
log[(t + dt)/t] log[t + dt] − log[t]
. 1 2 L

(S2) (S3)
[j]
For an observable scaling as a power-law with exponent where are Dj × Dj+1 matrices (D1 = DL+1 = 1),
Asj ,s′
δ ′ (O(t) = O0 tδ ), δO (t) = δ ′ is a constant. Generally, d = 2 is the physical on-site Hilbert space dimension,
′ j

we expect the observables to scale as power-laws at late and L is the system size. During the evolution, two-
times at the critical point p = pc , with δO approaching qubit unitary gates U (and one-qubit channels E) are
a constant at late t. For the classical point simulations, applied to the state, updating it according to ρ → U ρU †
we use a large value of dt = 500 to reduce the statistical (ρ → E[ρ]). During the U updates, two neighboring A
errors on effective exponents. tensors are contracted together forming a larger two-site
Fig. S7a shows the time-dependence of the total num- tensor (see Fig. S9), which can be converted back into
ber of active sites ⟨N (t)⟩ at the classical point for dif- two single-site tensors using a singular value decompo-
ferent values of p. Fig. S7b shows the time-dependence sition (SVD). We perform a truncated SVD after each
of the effective exponent of ⟨N (t)⟩ for different p. From U update, so that only the D largest singular values are
this we can clearly see how the effective exponent asymp- kept during each SVD (this results in Dj ≤ D; D is called
totes to a constant close to the DP value of Θ = 0.313686 the bond dimension). We keep the MPO normalized so
for p ≈ 0.394, but deviates from this constant at late that tr(ρ) = 1 after each SVD. We also keep the MPO
8

a b

c d

FIG. S7. Classical point observables and effective exponents. The time dependence of a the total number of active
sites, b the root-mean-squared distance, c the survival probability, and d the center-site density and their effective exponents at
the classical point θ = π of the FQCP. The observables were estimated using 105 Markov chain samples. The values of directed
percolation exponents are marked with horizontal dashed black lines.

a b

FIG. S8. Classical point effective exponent crossing. a The effective exponent of the total number of active sites ⟨N (t)⟩
versus p for fixed t. The DP value of the critical exponent Θ is marked with a black dashed line. The values were estimated
using 105 Markov chain samples. b A zoomed-in view of the effective exponent crossings, obtained from 1.12 × 107 samples.
The crossings of the t = 450 and t = 500 curves are marked with a red star and red dashed lines.

in mixed canonical form so that the orthogonality center 2. Analysis of simulation errors
is centered at one of the two sites where U is applied
before each U update. Performing a single layer of gate
and channel updates requires ∼ LD3 time. For the single
It is important to note that the approximate density
matrix ρ obtained after a truncated SVD is not neces-
seed initial state, performing updates at times 0, 1, . . . , t sarily Hermitian or positive semi-definite like a true den-
effectively involves performing layer updates on systems
of size L = 1, 5, . . . , 4t + 1. Therefore, performing a full
sity matrix. Unfortunately, it is in general difficult to
even check if a general MPO is positive semi-definite21–23 .
MPO dynamics simulation up to time t requires ∼ t2 D3 Moreover, quantifying the errors for MPO density matrix
time. evolution is not as straight-forward as for matrix prod-
9

a
a c
𝜌= 𝐴[1] 𝐴[𝐿] 𝐴[𝑖] 𝐸 𝐴[𝑖] ′

b 𝑈
SVD
𝐴[𝑖] 𝐴[𝑗] 𝐴[𝑖] ′ 𝐴[𝑗] ′

𝑈†
b
FIG. S9. MPO dynamics simulation. a The density ma-
trix ρ as a matrix product operator (MPO). b Update of the
MPO upon application of a two-qubit gate U , involves a sin-
gular value decomposition (SVD). c Update of the MPO upon
application of a single-qubit channel E.

uct state (MPS) pure-state evolution. In MPS dynam-


ics simulations, the sum of the discarded singular values
squared during a truncated SVD bounds the `2 -norm-
error of the pure state, which in turn bounds errors on
c
observables. However, in MPO dynamics simulations,
the same quantity only bounds the `2 -norm-error of the
mixed state, but does not tightly bound the errors on ob-
servables (they are more tightly bound by the `1 or trace
norm). In our MPO simulations, we find that errors on
observables can be many orders of magnitude larger than
the sum of the discarded singular values squared. Due to
this observation, we choose to not use an adaptive bond-
dimension technique, but instead choose a fixed bond
dimension D and examine how observables depend on
D.
FIG. S10. MPO dynamics simulation singular value de-
We empirically studied the errors in the MPO simula- composition errors. The a cumulative sum of the discarded
tion by looking at a few proxies for simulation inaccuracy. singular values squared ∆`2 , b cumulative sum of ⟨N (t)⟩ dif-
First, we considered the cumulative error due to singular ferences from truncated SVD, and c cumulative sum of P (t)
value truncations: differences from truncated SVD versus bond dimension D for
many values of t near the critical point at p = 0.3.
(k)
⎛ ∑j=D+1
χk
(sj )2 ⎞
∆`2 = ∑
⎝ ∑χk (s(k) )2 ⎠
(S4)
k l=1 l answer, but does provide a rough proxy for this quan-
where k indexes the SVDs performed during the simula- tity. Fig. S10 shows the cumulative SVD errors ∆`2 and
(k) (k)
tion and s1 ≥ . . . ≥ sχk are the ordered singular values
cumulative errors in total active sites ∆⟨N (t)⟩ and sur-
vival probability ∆P (t) for different bond dimensions at
p = 0.3.
obtained during the k-th SVD. The term in parenthe-
ses is a (normalized) sum of the squares of the singular
In addition to measuring the accumulated errors from
values discarded during the k-th SVD. Additionally, we
each SVD, we also measured the relative observable er-
ror ∣OD − OD=D0 ∣/∣OD=D0 ∣, where OD is the observable
also recorded the cumulative errors in observables due to
obtained from a finite D MPO simulation and D0 = 1024
SVD truncation:

∆O = ∑ ∣Ok − Ok′ ∣
is the largest bond dimension simulated. Fig. S11 shows
k the survival probability P (t) observable and its relative
errors versus bond dimension. Interestingly, the relative
where Ok is the observable computed before the k-th error does not monotonically decrease with D. This could
SVD truncation and Ok′ is the observable computed af- possibly be due to the non-monotonic dependence of ob-
ter. This quantity does not bound the inaccuracy of the servables with bond dimension (see Fig. S11b) or poten-
computed finite-D observable relative to the true D → ∞ tially due to numerical convergence issues in the MPO
10

simulations (though there is not evidence for this in the value, or optimal ω). Fig. S13c shows the p and effective
cumulative SVD errors shown in Fig. S10). exponent values at the crossings, as well as the optimal
BST extrapolation obtained from the accurate early-time
crossing data with the DP exponent values marked for
3. Simulation results and determination of critical point comparison. Fig. S13d shows the extrapolation error,
properties extrapolated pc , and extrapolated exponent values as a
function of ω for the extrapolation shown in Fig. S13c.
Note that we estimate the locations of the crossings by
In our simulations, we time-evolve the single seed ini-
tial state up to time t = 20. To avoid any finite system-
performing linear interpolation between the two curves;
size effects, we use a finite MPO of length L = 4t + 2 = 82
to ensure accurate crossing values, we perform MPO sim-
sites. We perform simulations at θ = 3π/4 for many val-
ulations at enough values of p so that each computed
crossing value pc (t) is between two simulated p values
ues of p and use bond dimensions up to D = 1024 (un-
spaced at most 2 × 10−4 apart.
less specified otherwise, all figures show θ = 3π/4 and
D = 1024 results). Our MPO code is written in Julia We apply the same procedure to the the mean-squared
(version 1.6.2)20 and uses the ITensor library (version displacement ⟨R2 (t)⟩, survival probability P (t), and
0.2.9)24 . center-site density ⟨n(r = 0, t)⟩. From these anal-
yses, we estimate the critical point locations pc ≈
0.3064, 0.3067, 0.304, 0.29, which suggests that 0.29 ≲ pc ≲
Fig. S12 shows the values of observables measured from
our MPO dynamics simulations versus time t for many
0.31. Most of the effective exponent values obtained
fixed p values. For each observable O(t), we also com-
puted an effective exponent (Eq. (S2) with dt = 2).
this way are close to the DP values, deviating by 0.92%,
0.28%, 4.4% relative error for ⟨N (t)⟩, ⟨R2 (t)⟩, and P (t),
To estimate the location of the critical point in our
respectively (see Table S1 for comparison of exponents
model, we follow a procedure utilized in previous numer-
with DP values). The small spread in pc estimates and
ical studies of non-equilibrium phase transitions25,26 . We
compute the effective exponents δO (t; p) for fixed t and
the closeness of the effective exponents to the known DP
many different values of p. We then identify the p = pc (t)
values provide good evidence that the FQCP has a DP
transition at pc ≈ 0.3 for θ = 3π/4. Notably, the center-
values at which δO (t + τ ; pc (t)) = δO (t; pc (t)), i.e., where
the δO (t + τ ; p) and δO (t; p) curves cross (we use τ = 2).
site density’s effective exponent appears to deviate sig-
Fig. S13a shows the effective exponent curves for ⟨N (t)⟩,
nificantly, about 31%, from the DP value. Some possible
with the (pc (t), δ⟨N (t)⟩ (t; pc (t))) crossing points marked
explanations for this deviation are: strong finite-time ef-
fects for this observable and strong finite bond-dimension
by black stars. These crossing points drift towards the
true critical point as t increases so that pc (t → ∞) =
effects (the center site in the 1D MPO is the most highly
entangled site and the site most affected by SVD trunca-
pc approaches the true critical decay probability and
δ⟨N (t)⟩ (t → ∞; pc (t → ∞)) = Θ approaches the true criti-
tion). Evidence that this discrepancy is due to short-time
effects is given in Fig. S15, which shows that the center-
cal exponent. However, for small finite t, there can be sig-
site density at the classical point also exhibits a critical
nificant finite-time deviations from the infinite-time limit
exponent far from the DP value at early times. From
results. Following prior work25–27 , we use a technique
Fig. S7d, we see that eventually, at late enough times,
known as Bulirsch-Stoer (BST) extrapolation28 to esti-
mate the t → ∞ limit results for pc (δO ) from the finite
the classical point center-site density does eventually ap-
pc (t) (δO (t; pc (t))) data. The BST technique involves
proach the expected DP power-law scaling.
We also performed MPO dynamics simulations for two
a free parameter ω, which roughly corresponds to the
modified version of the FQCP model: one with all con-
leading order correction to the infinite-time value (e.g.,
pc (t) ≈ pc + at−ω + ⋯). We determine the optimal ω value
trolled X rotations (instead of X and Y rotations) and
one with single-qubit amplitude damping channels in-
for the extrapolation by performing a fine scan over ω
values from 10−3 to 5 and finding the value that mini-
stead of probabilistic reset channels. Using the same
effective-exponent-crossing analysis described above, we
found similar evidence of a transition at pc ≈ 0.3 with
mizes an “extrapolation error” objective function. This
extrapolation error, used in Ref.26 , is defined as the sum
effective exponents consistent with directed percolation.
of the differences between BST extrapolated values ob-
These results suggest that the FQCP model is not fine-
tained from the original data set and modified data sets
tuned and that different unitary gates and dissipative
with one value removed. Fig. S13b shows the extrapola-
channels can also lead to the same directed percolation
tion error, extrapolated values, and optimal ω parame-
ters obtained from data sets containing the ⟨N (t)⟩ cross-
transition.
ings up to a time t. Due to the finite bond dimension
D used in the MPO numerics, the effective exponents
are less accurate at larger t (and smaller p), where cor- 4. Validation of methods on classical point
relations are the strongest. Because of this, the late-
time crossings are likely inaccurate, which can be seen in To validate our approach for locating the transition
abrupt changes in the BST extrapolation at late times and determining the critical exponents, we also per-
(e.g., abrupt jumps in extrapolation error, extrapolated formed the same MPO dynamics simulations and effec-
11

a b

c d

FIG. S11. MPO dynamics observable errors. a The survival probability P (t) versus bond dimension D for fixed times t at
p = 0.3. b A zoomed-in view of P (t = 14) versus D, highlighting the non-monotonic convergence of observables with increasing
bond dimension. c The relative error in P (t) versus D compared with the largest bond dimension D = 1024 result at fixed t,
on a log-linear scale. d The relative error in P (t) versus time t at fixed D, on a log-log scale.

a b

c d

FIG. S12. Observables and effective exponents. The time dependence of a the total number of active sites, b the root-
mean-squared distance, c the survival probability, and d the center-site density and their effective exponents, for a D = 1024
bond dimension MPO simulation.
12

a b

c d

FIG. S13. BST extrapolation of ⟨N (t)⟩. a The effective exponents of the observable versus p for different times t, with
crossings between t and t + 2 curves marked with black stars. b The optimal BST extrapolations using the crossings with
times up to t. c The p and effective exponent values at the marked crossings in a versus time t. The orange stars indicate
crossings that are likely inaccurate due to finite bond-dimension effects (in b, we see dramatic changes in the performance of
the BST extrapolation when using these crossings). d The optimal BST extrapolation data for p and the effective exponent
versus parameter ω using all of the accurate crossings indicated in c. Black dashed lines indicate the critical exponent values
for directed percolation and the red dotted lines indicate our best BST extrapolations.
13

Observable Exponent FQCP DP Relative error appears similar to “entanglement barriers” observed in
⟨N (t)⟩ Θ 0.3166 0.313686 0.92% prior studies of open quantum systems29 . Naively, this
⟨R2 (t)⟩ 2/z 1.2616 1.265226 0.28% suggests that one does not need a large bond dimension
P (t) −β/ν∥ −0.1525 −0.159464 4.4% MPO to accurately capture the dynamics of the system
⟨n(r = 0, t)⟩ Θ − 1/z −0.2186 −0.318927
at late times, even at the critical point. However, by
31%
considering the convergence of observables with bond di-
TABLE S1. Estimated FQCP critical exponents. For mension, we find that one does need a significant bond
four observables, which at the critical point scale as power- dimension (that increases with t) to obtain accurate sim-
laws in time ∼ tδ for exponent δ, we present our estimates for ulation results for p ≲ pc . This suggests that the MPO
the critical exponents for the Floquet quantum contact pro- entanglement entropy does not always directly coincide
cess (FQCP) at the quantum point θ = 3π/4. These estimates with classical simulation difficulty, as suggested in a re-
were obtained from the effective-exponent-crossing analysis cent study29 . Fig. S16b shows the half-cut Renyi mu-
described in the text, which used MPO dynamics simulation tual information I2 = S2 (ρA ) + S2 (ρB ) − S2 (ρ), where ρA
data. For comparison, we include the directed percolation (ρB ) is the reduced density matrix of the left-half (right-
(DP) values18,19 of the exponents as well as relative errors
half) of the system and S2 (ρ) = − log tr(ρ2 ). Unlike the
between the FQCP and DP values.
MPO entanglement entropy, the Renyi mutual informa-
tion grows with t for p ≲ pc ; it appears to more accu-
tive exponent crossing analysis at the classically solvable rately capture the presence of (classical and quantum)
point θ = π. As shown in Fig. S14, we found that the correlations in the system than the MPO entanglement
errors in the MPO simulation at the classical point were entropy. Fig. S16c shows the purity tr(ρ2 ) for the full
density matrix ρ. Note that for the single seed initial
state considered here, the density matrix has ≳ p overlap
much lower than at the quantum point (similar results
with the absorbing state (ρ ≈ p∣0⋯0⟩⟨0⋯0∣ + ⋯), which
were observed in MPO dynamics simulations performed
in Ref.6 ). Fig. S15 shows the effective exponent cross-
ings for ⟨N (t)⟩, ⟨R2 (t)⟩, P (t), and ⟨n(r = 0, t)⟩ ob- means that tr(ρ2 ) ≳ p2 . This prevents the purity from
tained from bond dimension D = 512 MPO dynamics going to zero in the active phase.
simulations. The relative errors of the effective expo-
nents from the DP values are 3.5%, 1.0%, 7.2%, 15.7%,
B. Open DMRG simulation of quasi-steady state
respectively. Like for the quantum point MPO simula-
tions, the ⟨n(r = 0, t)⟩ effective exponent deviates sub-
stantially from the true value due to strong finite t ef- As a complement to the MPO dynamics simulation,
fects. Moreover, the estimates of the critical point of we also implement an open density matrix renormaliza-
pc ≈ 0.3936, 0.3971, 0.3960, 0.3890 obtained from BST tion group (DMRG) calculation of the low-lying excited
extrapolation are within 0.2%, 0.7%, 0.4%, 1.4% of the states of the Floquet channel. Compared to the dynam-
pc = 0.3944 estimate obtained from the high-accuracy ical simulation, this gives access to very late times and
classical point numerics discussed above (see Fig. S8). can be advantageous in cases where there is an opera-
These results suggest that our procedure for estimating tor entanglement barrier to extrapolating from short to
critical points from short-time MPO dynamics simula- long times. The DMRG results are consistent with the
tions can accurately estimate the presence of the phase DP universality class when appropriately extrapolated to
transition and that the quantum point and classical point infinite system size.
both exhibit DP critical scaling.

1. Details of the method


5. Dynamics of entanglement entropies and other
quantities In general, the Floquet evolution superoperator E =
E(θ, p) has eignvalues eεα given by E[ρα ] = eεα [ρα ]. As-
In our MPO simulations, we also compute quantities, suming there is no dark steady state, these eigenvalues
such as entropies, that cannot be measured (or are dif- can be sorted as 0 = ε0 > Reε1 ⩾ Reε2 ⩾ ⋯. For the FQCP
ficult to measure) experimentally. Fig. S16a shows the model, the absorbing state ρ0 = ∣0⋯0⟩⟨0⋯0∣ is the unique
half-cut MPO entanglement entropy S = − ∑m pm log pm steady state in finite size system, thus the key quantity
where pm = s2m / ∑k s2k and sm are the singular values of is the relaxation time from the longest-lived active state
the A[j] tensor (shaped as a matrix with the columns cor- to the absorbing inactive state, which here is given by
responding to the bond between sites j and j + 1). Note τ = − Reε
1
1
corresponding to the lifetime of the slowest de-
that this quantity, despite its name, does not measure cay mode ρ1 . Finite-size scaling implies that, at the crit-
entanglement but rather computes the entropy present ical point, the relaxation time grows with systems size as
in the singular value spectra of the tensors connecting τ ∼ Lz , which hence provides an approach to determine
sites j and j + 1. Surprisingly, we find that, after a pc and z. Since the relaxation time in the active phase
brief increase in the first time step, the MPO entan- is in general exponentially large, it is difficult to calcu-
glement entropy quickly decays to zero, for all p. This late the relaxation time directly through time-evolution
14

FIG. S14. MPO dynamics errors at classical point. The cumulative sum of a the discarded singular values squared
∆`2 , b ⟨N (t)⟩ differences from truncated SVD, and c P (t) differences from truncated SVD versus bond dimension D for many
values of t, on a log-linear scale. The relative error in d ⟨N (t)⟩ and e P (t) versus D compared with the largest bond dimension
D = 512 result at fixed t, on a log-linear scale. The results are at the classical point θ = π at p = 0.394 near the critical point.

methods. As an alternative, we use the density-matrix notation, the Floquet evolution superoperator E, which
renormalization group (DMRG) scheme to obtain a nu- implements one Floquet period of time evolution, can be
merical approximation of the slowest decay mode ρ1 of written as a matrix-product (super) operator acting on
the Floquet evolution superoperator E 30–32 . ∣ρ⟫:
We first consider a vectorization of the density matrix
ρ → ∣ρ⟫ using the Choi isomorphism ∣ψ⟩⟨φ∣ → ∣ψ ⊗ φ⟫. E = ∑ Mµ[1]
1 ν1
⋯Mµ[L]
L νL
∣µ1 ⋯µL ⟫⟪ν1 ⋯νL ∣, (S6)
The density matrix can be represented as an MPO or an {µi ,νi }
MPS with squared physical dimension, which is compat-
ible with DMRG (see Fig. S17 for a graphical represen- where each M [i] is a d2 ×d2 ×DO,i ×DO,i+1 tensor, DO,1 =
tation): DO,L+1 = 1, and DO,i ≤ DO where DO is the operator
bond dimension.
∣ρ⟫ = ∑ A[1]
µ1 . . . AµL ∣µ1 . . . µL ⟫.
[L]
(S5) To target the slowest decaying mode, ρ1 , we leverage
{µi } that the trace-preserving nature of E implies a model-
independent left steady state: ⟪I∣E = ⟪I∣, where ∣I⟫ is
Here, each A[i] is a d2 × Di × Di+1 tensor, µi ∈ {1 . . . d2 } the maximally mixed state, i.e., ⟪I∣ is the left eigenvector
labels a basis of physical states for the vectorized density corresponding to steady state ∣ρ0 ⟫. We use this fact to
matrix, i = 1 . . . L label sites of the system, D1 = DL+1 = 1 construct the modified evolution operator E ′ = E −w∣I⟫⟪I∣,
and Di ≤ D where D is the bond dimension. In this where w > 0 penalizes the steady (absorbing) state and
15

a b

c d

FIG. S15. Determination of critical point scaling at classical point. The crossings of the effective exponents for the a
the total number of active sites, b the root-mean-squared distance, c the survival probability, and d the center-site density for
a D = 512 bond dimension MPO simulation at the classical point θ = π. The black dashed lines correspond to DP exponent
values and the red dashed lines correspond to BST extrapolations based on the crossings.

causes ρ1 to become the “quasi-steady state” of E ′ , or size critical point pc (L) and critical exponent z(L) satis-
eigenvector with eigenvalue closest to 1. Using the modi- fying RL (pc (L)) = RL+2 (pc (L)) ≡ z(L). Both pc (L) and
fied Floquet operator E ′ , we find an MPS approximation z(L) form finite-size sequences, from which we perform
of ρ1 by iteratively solving for the eigenvalue with largest BST extrapolation to estimate pc = pc (∞) and z = z(∞).
real part of the local effective evolution operator Ei′ ob- The simulation used to extract the critical point infor-
tained by contracting all indices of ⟪ρ∣E ′ ∣ρ⟫ except for mation is performed from L = 11 to L = 39 with maxi-
the indices on site i (see Fig. S17 for graphical represen- mum bond dimension Dmax = 324 for the classical case
tation). (In practice, we take w = 1 and have DO = 256 and from L = 11 to L = 31 for the quantum case with
for the FQCP model.) Dmax = 260. As shown in Fig.S18(c,d), for the classi-
cal point θ = π, we obtain pc ≈ 0.3945, which is con-
sistent with the high-accuracy result pc = 0.3944 men-
2. Critical point scaling of relaxation time tioned above, and z ≈ 1.572 with 0.5% off the DP value
z = 1.5807. For the quantum case θ = 43 π, the estimated
We perform DMRG simulations at both the classical critical point is pc ≈ 0.3091 with z ≈ 1.515, which has
point θ = π and quantum point θ = 3π/4. The inverse of 4.15% deviation compared to the DP value.
the relaxation time, τ −1 , which we refer to as the dissi-
pative gap, is shown in Fig.S18 (a) and (b). To extract
the critical point information, we consider the finite-size We note that that the errors in the DMRG calcula-
scaling ansatz τ (p, L) = L−z F(∣p − pc ∣L1/ν⊥ ), where F is tion, for example due to bond-dimension truncation or
a scaling function. This scaling form implies that the due to the finite tolerance of the non-Hermitian eigen-
quantity25,26 solver used in DMRG, propagate through to errors in
the finite-size extrapolation for pc and hence z in a man-
log [τ (p, L)/τ (p, L − 2)]
RL (p) ≡
ner that is difficult to reliably estimate. We expect that
log [(L − 2)/L]
, (S7) such errors account for the small percent-level discrep-
ancies between the extracted z value and the DP values,
converges to z at p = pc as L → ∞. By considering three and view the DMRG results as consistent within error
consecutive sizes L − 2, L, L + 2, one can define the finite- bars of DP scaling.
16

FIG. S16. MPO dynamics entropies. a Half-cut MPO


entanglement entropy S, b half-cut Renyi mutual information
I2 , and c purity tr(ρ2 ) of the full density matrix versus time
t for different p. Results for θ = 3π/4.

∣ρ⟫ = A A A A A

Ā Ā Ā Ā

Ei′ = M M M M M

A A A A
FIG. S17. DMRG simulation. (upper) Vectorized density
matrix as a matrix product state (MPS); (lower) Modified
local Floquet evolution opertor Ei′ .
17

=
(a) L = 11 L = 23 (c)0.450 (e)
0.15 L = 13 L = 25
102
L = 15 L = 27 0.425
L = 17 L = 29 100
L = 19 L = 31 0.400 =

BST error
0.10 = 3 /4
L = 21 10 2

pc(L)
0.375 pc( , = ) = 0.3945
1/

pc( , = 3 /4) = 0.3091 10 4


0.05 0.350
0.325 10 6 pc
0.00 z
0.1 0.2 = 30.3/4 0.4 10 20 30 40 0 1 2 3 4 5
(b) (d) 1.6 (f)
0.15 1.5 104
102
1.4

BST error
0.10 100
1.3
z(L)
1/

10 2
0.05 1.2
DP : z = 1.5807
10 4
1.1 z( , = ) = 1.572 10 6
0.00 z( , = 3 /4) = 1.515
1.0
0.1 0.2 0.3 0.4 10 20 30 40 0 1 2 3 4 5
p L

FIG. S18. Determination of critical point scaling of relaxation time. Inverse of relaxation time 1/τ versus p at the (a)
classical and (b) quantum points. (c) The finite-size critical point pc (L) versus L. (d) The finite-size critical exponent z(L)
versus L. The BST error from the extrapolation of pc (L) and z(L) at the (e) classical and (f) quantum points.

[1] J. M. Deutsch, Quantum statistical mechanics in a closed Rev. Research 2, 033284 (2020).
system, Phys. Rev. A 43, 2046 (1991). [11] S. Sugiura, T. Kuwahara, and K. Saito, Many-body scar
[2] M. Srednicki, Chaos and quantum thermalization, Phys. state intrinsic to periodically driven system, Phys. Rev.
Rev. E 50, 888 (1994). Research 3, L012010 (2021).
[3] M. Rigol, V. Dunjko, and M. Olshanii, Thermalization [12] B. Skinner, J. Ruhman, and A. Nahum, Measurement-
and its mechanism for generic isolated quantum systems, induced phase transitions in the dynamics of entangle-
Nature 452, 854 (2008). ment, Phys. Rev. X 9, 031009 (2019).
[4] M. Marcuzzi, M. Buchhold, S. Diehl, and I. Lesanovsky, [13] A. Zabalo, M. J. Gullans, J. H. Wilson, S. Gopalakrish-
Absorbing state phase transition with competing quan- nan, D. A. Huse, and J. H. Pixley, Critical properties of
tum and classical fluctuations, Phys. Rev. Lett. 116, the measurement-induced transition in random quantum
245701 (2016). circuits, Phys. Rev. B 101, 060301 (2020).
[5] F. Carollo, E. Gillman, H. Weimer, and I. Lesanovsky, [14] A. C. Potter and R. Vasseur, Entanglement dynamics in
Critical behavior of the quantum contact process in one hybrid quantum circuits, arXiv:2111.08018 (2021).
dimension, Phys. Rev. Lett. 123, 100604 (2019). [15] K. Temme, S. Bravyi, and J. M. Gambetta, Error mitiga-
[6] E. Gillman, F. Carollo, and I. Lesanovsky, Numerical tion for short-depth quantum circuits, Phys. Rev. Lett.
simulation of critical dissipative non-equilibrium quan- 119, 180509 (2017).
tum systems with an absorbing state, New J. Phys. 21, [16] Y. Li and S. C. Benjamin, Efficient variational quantum
093064 (2019). simulator incorporating active error minimization, Phys.
[7] M. Jo, J. Lee, K. Choi, and B. Kahng, Absorbing phase Rev. X 7, 021050 (2017).
transition with a continuously varying exponent in a [17] C. Ryan-Anderson, N. C. Brown, M. S. Allman, B. Arkin,
quantum contact process: A neural network approach, G. Asa-Attuah, C. Baldwin, J. Berg, J. G. Bohnet,
Phys. Rev. Research 3, 013238 (2021). S. Braxton, N. Burdick, J. P. Campora, A. Chernoguzov,
[8] Y. Y. Atas, E. Bogomolny, O. Giraud, and G. Roux, J. Esposito, B. Evans, D. Francois, J. P. Gaebler, T. M.
Distribution of the ratio of consecutive level spacings in Gatterman, J. Gerber, K. Gilmore, D. Gresh, A. Hall,
random matrix ensembles, Phys. Rev. Lett. 110, 084101 A. Hankin, J. Hostetter, D. Lucchetti, K. Mayer, J. My-
(2013). ers, B. Neyenhuis, J. Santiago, J. Sedlacek, T. Skripka,
[9] L. D’Alessio and M. Rigol, Long-time behavior of isolated A. Slattery, R. P. Stutz, J. Tait, R. Tobey, G. Vittorini,
periodically driven interacting lattice systems, Phys. J. Walker, and D. Hayes, Implementing fault-tolerant en-
Rev. X 4, 041048 (2014). tangling gates on the five-qubit code and the color code,
[10] K. Mizuta, K. Takasan, and N. Kawakami, Exact floquet arXiv:2208.01863 (2022).
quantum many-body scars under rydberg blockade, Phys. [18] I. Jensen, Low-density series expansions for directed per-
18

colation: I. a new efficient algorithm with applications to (2008).


the square lattice, J. Phys. A 32, 5233 (1999). [26] J. C. M. Filho and R. Dickman, Conserved directed per-
[19] H. Hinrichsen, Non-equilibrium critical phenomena and colation: exact quasistationary distribution of small sys-
phase transitions into absorbing states, Adv. Phys. 49, tems and monte carlo simulations, J. Stat. Mech.: Theory
815 (2000). Exp. 2011 (05), P05029.
[20] J. Bezanson, A. Edelman, S. Karpinski, and V. B. Shah, [27] E. Carlon, M. Henkel, and U. Schollwöck, Density matrix
Julia: A fresh approach to numerical computing, SIAM renormalization group and reaction-diffusion processes,
Review 59, 65 (2017). Eur. Phys. J. B 12, 99 (1999).
[21] M. Kliesch, D. Gross, and J. Eisert, Matrix-product oper- [28] M. Henkel and G. Schutz, Finite-lattice extrapolation al-
ators and states: Np-hardness and undecidability, Phys. gorithms, J. Phys. A 21, 2617 (1988).
Rev. Lett. 113, 160503 (2014). [29] K. Noh, L. Jiang, and B. Fefferman, Efficient classical
[22] A. H. Werner, D. Jaschke, P. Silvi, M. Kliesch, simulation of noisy random quantum circuits in one di-
T. Calarco, J. Eisert, and S. Montangero, Positive tensor mension, Quantum 4, 318 (2020).
network approach for simulating open quantum many- [30] J. Cui, J. I. Cirac, and M. C. Bañuls, Variational ma-
body systems, Phys. Rev. Lett. 116, 237201 (2016). trix product operators for the steady state of dissipative
[23] C. D. White, M. Zaletel, R. S. K. Mong, and G. Refael, quantum systems, Phys. Rev. Lett. 114, 220601 (2015).
Quantum dynamics of thermalizing systems, Phys. Rev. [31] E. Mascarenhas, H. Flayac, and V. Savona, Matrix-
B 97, 035127 (2018). product-operator approach to the nonequilibrium steady
[24] M. Fishman, S. R. White, and E. M. Stoudenmire, The state of driven-dissipative quantum arrays, Phys. Rev. A
ITensor Software Library for Tensor Network Calcula- 92, 022116 (2015).
tions, SciPost Phys. Codebases , 4 (2022). [32] Z. Cheng and A. C. Potter, A matrix product opera-
[25] R. Dickman, Absorbing-state phase transitions: Exact tor approach to non-equilibrium floquet steady states,
solutions of small systems, Phys. Rev. E 77, 030102 arXiv:2206.07740 (2022).

You might also like