Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2310.03078v2 [cond-mat.dis-nn] 05 Jan 2024

Boundary transfer matrix spectrum of measurement-induced transitions

Abhishek Kumar Department of Physics, University of Massachusetts, Amherst, Massachusetts 01003, USA    Kemal Aziz Department of Physics and Astronomy, Center for Materials Theory, Rutgers University, Piscataway, NJ 08854, USA    Ahana Chakraborty Department of Physics and Astronomy, Center for Materials Theory, Rutgers University, Piscataway, NJ 08854, USA    Andreas W. W. Ludwig Department of Physics, University of California, Santa Barbara, California 93106, USA    Sarang Gopalakrishnan Department of Electrical and Computer Engineering, Princeton University, Princeton, NJ 08544, USA    J. H. Pixley Department of Physics and Astronomy, Center for Materials Theory, Rutgers University, Piscataway, NJ 08854, USA Center for Computational Quantum Physics, Flatiron Institute, 162 5th Avenue, New York, NY 10010, USA    Romain Vasseur rvasseur@umass.edu Department of Physics, University of Massachusetts, Amherst, Massachusetts 01003, USA
(January 5, 2024)
Abstract

Measurement-induced phase transitions (MIPTs) are known to be described by non-unitary conformal field theories (CFTs) whose precise nature remains unknown. Most physical quantities of interest, such as the entanglement features of quantum trajectories, are described by boundary observables in this CFT. We introduce a transfer matrix approach to study the boundary spectrum of this field theory, and consider a variety of boundary conditions. We apply this approach numerically to monitored Haar and Clifford circuits, and to the measurement-only Ising model where the boundary scaling dimensions can be derived analytically. Our transfer matrix approach provides a systematic numerical tool to study the spectrum of MIPTs.

preprint: APS/123-QED

I Introduction

Repeated measurements can drive phase transitions in the entanglement structure of quantum trajectories of many-body quantum systems Li et al. (2018, 2019); Skinner et al. (2019); Chan et al. (2019); Choi et al. (2020); Potter and Vasseur (2022); Fisher et al. (2023). The discovery of such measurement-induced phase transitions (MIPTs) has attracted a lot of attention recently Cao et al. (2019); Szyniszewski et al. (2019); Choi et al. (2020); Lunt and Pal (2020); Goto and Danshita (2020); Tang and Zhu (2020); Iaconis et al. (2020); Turkeshi et al. (2020); Zhang et al. (2020); Szyniszewski et al. (2020); Fuji and Ashida (2020); Rossini and Vicari (2020); Vijay (2020); Turkeshi et al. (2021); Chen (2021); Li and Fisher (2021a); Lavasani et al. (2021a); Van Regemortel et al. (2021); Lunt et al. (2021); Alberton et al. (2021); Nahum et al. (2021); Han and Chen (2022); Sierant et al. (2022a); Sharma et al. (2022); Feng et al. (2022); Iadecola et al. (2022); Buchhold et al. (2022); Sierant and Turkeshi (2023a); O’Dea et al. (2022); Piroli et al. (2023); Sierant and Turkeshi (2023b); LeMaire et al. (2023); Ravindranath et al. (2023a); Jian et al. (2023); Fava et al. (2023); Ravindranath et al. (2023b); Roser et al. (2023); Coppola et al. (2022); Tirrito et al. (2023), mostly in the context of monitored quantum circuits, consisting of entangling unitary gates and disentangling local measurement operators. Focusing on the properties of quantum states conditional on the measurement outcomes, the unitary dynamics results in the scrambling of quantum information and volume-law entanglement scaling, whereas increasing the rate of local measurements can eventually lead to area-law scaling. This transition can also be equivalently interpreted as a purification transition Gullans and Huse (2020a), a quantum coding transition Choi et al. (2020); Li and Fisher (2021b); Lovas et al. (2023), or a learning transition Agrawal et al. (2022); Barratt et al. (2022a, b); Majidy et al. (2023); Li et al. (2023a); Tikhanovskaya et al. (2023); Ippoliti and Khemani (2023) quantifying how much information the observer learns from the measurement records.

Given these diverse interpretations and applications, a natural question is to understand the critical behavior of this transition. A crucial step in that direction is provided by exact mappings onto replica statistical mechanical models where the transition is interpreted as an ordering transition from ferromagnetic (volume law) to paramagnetic (area law) phases Bao et al. (2020); Jian et al. (2020); Nahum et al. (2021); Li et al. (2021a, 2023b) – see also Hayden et al. (2016); Vasseur et al. (2019); Zhou and Nahum (2019) for earlier results on random tensor networks and random unitary circuits. In turn, this statistical mechanics mapping can be used to formulate effective field theory descriptions of MIPTs Jian et al. (2020); Nahum et al. (2021); Nahum and Joerg Wiese (2023), which remain to be fully understood. A key prediction of the statistical mechanics mapping is the emergence of conformal invariance at the critical point, which was first observed numerically in monitored Clifford circuits Li et al. (2021b). More precisely, MIPTs in 1+1d generic monitored quantum systems are described by non-unitary conformal field theories (CFTs) with central charge c=0𝑐0c=0italic_c = 0, also known as logarithmic CFTs Gurarie (1993); Gurarie and Ludwig (2005); Cardy (2013) (or log-CFTs for short). Consequently, conformal invariance plays a crucial role in precise characterization of the nature of MIPTs Li et al. (2021b); Zabalo et al. (2020); Li et al. (2021a); Sierant et al. (2022b); Block et al. (2022); Zabalo et al. (2022); Tikhanovskaya et al. (2023).

While a full analytic understanding of the CFTs underlying MIPTs remains out of reach at the moment, it is possible to utilize conformal invariance to study their properties numerically, at least in one dimension. In recent work Zabalo et al. (2022), it was argued that the quantum evolution with fixed measurement outcomes can be interpreted as a disordered transfer matrix which can be used to extract critical properties using standard CFT tools. This approach was used to extract new universal properties, but also provided numerical estimates of various bulk scaling dimensions accurate enough to distinguish MIPTs in generic monitored circuits (sampled with Haar measure) from those in Clifford monitored circuits.

This transfer matrix approach relies on using periodic boundary conditions. It probes the bulk properties of the underlying CFT—including the effective central charge, the order parameter, and the energy operator scaling dimensions—by putting it on an effectively infinitely long cylinder, and applying a finite size scaling analysis in the circumference. However, many physical quantities of interest are in fact boundary observables in the statistical mechanics model and in the CFT. This is because of the nature of the corresponding statistical mechanics model, which is defined on the geometry of the circuit: any physical quantity (including entanglement) computed at a given time t𝑡titalic_t in the monitored quantum circuit will map onto a statistical mechanics observable defined at the top boundary of a two-dimensional lattice (the space-time of the circuit). For example, the von Neumann entanglement entropy of an interval of size LAsubscript𝐿𝐴L_{A}italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT scales logarithmically with LAsubscript𝐿𝐴L_{A}italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT at criticality (at sufficiently long times)

SAγlnLA.similar-tosubscript𝑆𝐴𝛾subscript𝐿𝐴S_{A}\sim\gamma\ln L_{A}.italic_S start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∼ italic_γ roman_ln italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT . (1)

While in analogy with the entanglement structure of the groundstate of translationally invariant, i.e. non-random CFTs in 1+1d one could naively expect γ𝛾\gammaitalic_γ to be related to the central charge of the underlying CFT Calabrese and Cardy (2009), this is incorrect. Instead, the universal coefficient γ𝛾\gammaitalic_γ is related to a boundary scaling dimension – the scaling dimension of a so-called boundary condition changing (BCC) operator Jian et al. (2020); Li et al. (2021b), using terminology from boundary CFT (BCFT) Cardy (1984a, 1989). Note that, while as already mentioned the actual central charge is c=0𝑐0c=0italic_c = 0, the quantity playing instead a corresponding role in log-CFTs, including those discussed in this paper, is the so-called effective central charge ceffsubscript𝑐effc_{\rm eff}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, defined in Eqs. (4,5) below.

In this work, we introduce a boundary transfer matrix approach to study the BCFT data in various families of monitored circuits, which are believed to undergo MIPTs in distinct universality classes. We find that the boundary spectra for distinct universality classes are different. In addition, we numerically evaluate some new boundary exponents that had not yet been computed using different approaches.

The paper is organized as follows. In section (II) we discuss the boundary transfer matrix approach to extract the boundary spectrum of MIPTs in monitored 1+1d quantum systems. In section (III) we benchmark this approach using the measurement-only Ising model by comparing our numerical results to analytic predictions. In section (IV) and (V) we compute the boundary spectra of dual Clifford and dual Haar monitored random circuits, respectively. Finally we conclude in section (VI) by summarizing our results and discussing their broader implications.

II Boundary Transfer Matrix

Refer to caption
Figure 1: Boundary condition for monitored circuits: Implementation of three different boundary conditions for monitored circuits undergoing hybrid dynamics ((a𝑎aitalic_a)–(c𝑐citalic_c)). It consists of two qubit random unitary gates (blue rectangles) and σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT measurement operations (shown with red circles). All single qubit measurement operations occur at a given space-time location with a probability p𝑝pitalic_p. The initial state |ψ0ketsubscript𝜓0|\psi_{0}\rangle| italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ is evolved under hybrid dynamics with boundary ends set to be; (a) open or free, (b) dissipation (shown as purple diamond) at right end and other end is left free, and (c) entangled system-ancilla qubits at both ends. In particular we model dissipation by introducing either dephasing along zlimit-from𝑧z-italic_z -axis or by using a maximally mixed depolarizing channel. For the entangled system-ancilla setup we replace the boundary system qubit with a bell pair in each time period where the ancilla set comprise of one qubit from the Bell pairs (denoted by filled black circle) and all the discarded system qubits. The transfer matrix operator which encode the action of unitaries and measurements at the j𝑗jitalic_jth time step is given by 𝒯jsubscript𝒯𝑗\mathcal{T}_{j}caligraphic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

MIPTs are most well studied in hybrid circuits that consist of local projective measurements that are interspersed between two-site random unitary gates arranged in a brick-work pattern (as shown in Fig 1). The measurement probability p𝑝pitalic_p at a given space-time location is used to drive the transition. The resulting phases exhibit distinct steady state entanglement structure, conditional on measurement outcomes. At low p𝑝pitalic_p, the dynamics due to entangling unitary gates dominate which in turn results in a subsystem entanglement entropy that scales with the subsystem volume (volume-law), whereas at large p𝑝pitalic_p the local measurements effectively “collapse” the many-body wavefunction (area-law scaling). At the critical point p=pc𝑝subscript𝑝𝑐p=p_{c}italic_p = italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the entanglement scales logarithmically with the subsystem size following Eq. (1) and conformal invariance emerges.

In this work, we probe the boundary conformal properties (BCFT spectrum) of MIPTs by using a boundary transfer matrix, where different boundary conditions can be used to probe the scaling dimension of different BCC operators (i.e. boundary “observables”). This generalizes the bulk transfer matrix study of Ref. Zabalo et al., 2022.

We will apply this approach numerically to monitored circuits that feature Haar and Clifford dynamics, and to the measurement-only Ising model where the boundary scaling dimensions can be derived analytically. The Haar and Clifford circuits are realized by drawing the two-qubit unitary gates from the Haar distribution and the finite Clifford group, respectively. However, the transfer matrix spectrum extracted for random circuits with gates drawn from these groups have large error bars due to the non-universal anisotropy factor (α𝛼\alphaitalic_α) from the asymmetry in the circuit space and time direction Zabalo et al. (2022). This is improved by restricting the circuit dynamics to a smaller gate set called dual unitary gates Bertini et al. (2019); Gopalakrishnan and Lamacraft (2019); Piroli et al. (2020); Claeys et al. (2022), which are not expected to change the universality class Zabalo et al. (2022, 2020), but have anisotropy factor α=1𝛼1\alpha=1italic_α = 1 since the gates are unitary both in space and time direction.

As we detail below, we will focus on the following three boundary conditions at the two boundary ends of the infinite strip circuit geometry; (i) free boundary conditions, (ii) dissipator boundary using either dephasing or depolarizing channels, and (iii) “cyclic/swap” boundary conditions by probing the entanglement properties of boundary ancilla qubits  (see Fig 1). This last boundary condition (iii), however, is out of reach with Haar random gates as the number of ancilla qubits must grow with time. In spite of this limitation in the special case (iii), our analysis allows us to extract various boundary scaling dimensions of various MIPTs, and to further test the emergence of conformal invariance in monitored systems.

II.1 Transfer matrix

We consider monitored quantum circuits with open spatial boundary conditions, using various boundary conditions depicted in Fig 1. To begin, we consider a fixed set of space-time coordinates for the action of measurement operators and unitary gates. The dynamics is then described by the quantum channel

𝒩t(ρ)=𝐦K𝐦ρK𝐦,subscript𝒩𝑡𝜌subscript𝐦subscript𝐾𝐦𝜌superscriptsubscript𝐾𝐦\mathcal{N}_{t}(\rho)=\sum_{\bf{m}}K_{\bf{m}}\rho K_{\bf{m}}^{\dagger},caligraphic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_ρ ) = ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_ρ italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , (2)

where ρ𝜌\rhoitalic_ρ is system’s initial density matrix and K𝐦=KtmtKt1mt1K1m1subscript𝐾𝐦superscriptsubscript𝐾𝑡subscript𝑚𝑡superscriptsubscript𝐾𝑡1subscript𝑚𝑡1superscriptsubscript𝐾1subscript𝑚1K_{\bf{m}}=K_{t}^{m_{t}}K_{t-1}^{m_{t-1}}\dots K_{1}^{m_{1}}italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = italic_K start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_K start_POSTSUBSCRIPT italic_t - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_t - 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT … italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT where Ksms=PmssUssuperscriptsubscript𝐾𝑠subscript𝑚𝑠superscriptsubscript𝑃subscript𝑚𝑠𝑠subscript𝑈𝑠K_{s}^{m_{s}}=P_{m_{s}}^{s}U_{s}italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = italic_P start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT consists of the random unitary operations Ussubscript𝑈𝑠U_{s}italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT (possibly trivial for measurement-only dynamics) and random projectors Pmsssuperscriptsubscript𝑃subscript𝑚𝑠𝑠P_{m_{s}}^{s}italic_P start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT onto the measurement outcomes mssubscript𝑚𝑠m_{s}italic_m start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Each term (K𝐦ρK𝐦subscript𝐾𝐦𝜌superscriptsubscript𝐾𝐦K_{\bf{m}}\rho K_{\bf{m}}^{\dagger}italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_ρ italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT) in the sum of Eq. 2 represents a quantum trajectory which occurs with the Born probability p𝐦subscript𝑝𝐦p_{\mathbf{m}}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT (=Tr(K𝐦ρK𝐦)absentTrsubscriptK𝐦𝜌superscriptsubscriptK𝐦=\rm{Tr}(K_{\bf{m}}\rho K_{\bf{m}}^{\dagger})= roman_Tr ( roman_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_ρ roman_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT )) for the measurement outcomes 𝐦𝐦\bf{m}bold_m. The trajectories in the channel form an ensemble of statistical mechanical models with inherent space-time randomness coming from the measurement outcomes. Following Ref. Zabalo et al. (2022), we introduce the transfer matrix 𝒯j=K2jm2jK2j1m2j1subscript𝒯𝑗superscriptsubscript𝐾2𝑗subscript𝑚2𝑗superscriptsubscript𝐾2𝑗1subscript𝑚2𝑗1\mathcal{T}_{j}=K_{2j}^{m_{2j}}K_{2j-1}^{m_{2j-1}}caligraphic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_K start_POSTSUBSCRIPT 2 italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT 2 italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_K start_POSTSUBSCRIPT 2 italic_j - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT 2 italic_j - 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (shown in Fig 1) for the unitary-measurement dynamics which describes evolution for a single time period (maps ρ𝒯jρ𝒯j𝜌subscript𝒯𝑗𝜌superscriptsubscript𝒯𝑗\rho\to\mathcal{T}_{j}\rho\mathcal{T}_{j}^{\dagger}italic_ρ → caligraphic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ρ caligraphic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT). Now at late times the singular values σi𝐦subscriptsuperscript𝜎𝐦𝑖\sigma^{\bf{m}}_{i}italic_σ start_POSTSUPERSCRIPT bold_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (where (σi𝐦)2superscriptsubscriptsuperscript𝜎𝐦𝑖2(\sigma^{\bf{m}}_{i})^{2}( italic_σ start_POSTSUPERSCRIPT bold_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT are eigenvalues of K𝐦K𝐦subscript𝐾𝐦superscriptsubscript𝐾𝐦K_{\bf{m}}K_{\bf{m}}^{\dagger}italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT) of K𝐦subscript𝐾𝐦K_{\bf{m}}italic_K start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT decay exponentially σi𝐦=eλi𝐦𝐭subscriptsuperscript𝜎𝐦𝑖superscript𝑒superscriptsubscript𝜆𝑖subscript𝐦𝐭\sigma^{\bf{m}}_{i}=e^{\lambda_{i}^{\bf{m}_{t}}}italic_σ start_POSTSUPERSCRIPT bold_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_m start_POSTSUBSCRIPT bold_t end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT (where λi𝐦𝐭<0superscriptsubscript𝜆𝑖subscript𝐦𝐭0\lambda_{i}^{\bf{m}_{t}}<0italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_m start_POSTSUBSCRIPT bold_t end_POSTSUBSCRIPT end_POSTSUPERSCRIPT < 0) as the state purifies under evolution of maximally mixed density matrix (where (σi𝐦)2superscriptsubscriptsuperscript𝜎𝐦𝑖2(\sigma^{\bf{m}}_{i})^{2}( italic_σ start_POSTSUPERSCRIPT bold_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT are its eigenvalues). The trajectory averages of λi𝐦𝐭superscriptsubscript𝜆𝑖subscript𝐦𝐭\lambda_{i}^{\bf{m}_{t}}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_m start_POSTSUBSCRIPT bold_t end_POSTSUBSCRIPT end_POSTSUPERSCRIPT give the values of the Lyapunov exponents λ0,λ1,subscript𝜆0subscript𝜆1\lambda_{0},\lambda_{1},\dotsitalic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … (in descending order) of the transfer matrix. The leading Lyapunov exponent is related to the average free energy of the statistical mechanical model up to a factor of time, i.e., F=λ0t𝐹subscript𝜆0𝑡F=-\lambda_{0}titalic_F = - italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t. Interestingly, this is equivalent Zabalo et al. (2022) to the Shannon entropy of the measurement record, where

F=𝐦p𝐦lnp𝐦.𝐹subscript𝐦subscript𝑝𝐦subscript𝑝𝐦F=-\sum_{\bf{m}}p_{\bf{m}}\ln p_{\bf{m}}.italic_F = - ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT roman_ln italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT . (3)

At criticality, the scaling of this averaged free energy with system size L𝐿Litalic_L is dictated by conformal invariance. This can be seen by introducing the replicated partition functions Z¯k=𝐦(p𝐦p𝐦k)subscript¯𝑍𝑘subscript𝐦subscript𝑝𝐦superscriptsubscript𝑝𝐦𝑘\bar{Z}_{k}=\sum_{\bf{m}}(p_{\bf{m}}p_{\bf{m}}^{k})over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) from which the averaged free energy can be obtained as F=limk0dFkdk𝐹subscript𝑘0𝑑subscript𝐹𝑘𝑑𝑘F=\lim_{k\to 0}\frac{dF_{k}}{dk}italic_F = roman_lim start_POSTSUBSCRIPT italic_k → 0 end_POSTSUBSCRIPT divide start_ARG italic_d italic_F start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_k end_ARG in the replica limit k0𝑘0k\to 0italic_k → 0 where Fksubscript𝐹𝑘F_{k}italic_F start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT( =lnZ¯kabsentsubscript¯𝑍𝑘=-\ln{\bar{Z}_{k}}= - roman_ln over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT) is the replicated free energy. For any finite number of replicas k𝑘kitalic_k, Fksubscript𝐹𝑘F_{k}italic_F start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the free energy of a statistical mechanics model which exhibits (for k𝑘kitalic_k small enough) a 2nd order phase transition with emerging conformal invariance Bao et al. (2020); Jian et al. (2020). In the replica limit k0𝑘0k\to 0italic_k → 0, this transition coincides with the MIPT. To find the free energy density we have to take into account the space-time area (A𝐴Aitalic_A) where A=αLt𝐴𝛼𝐿𝑡A=\alpha Ltitalic_A = italic_α italic_L italic_t and α𝛼\alphaitalic_α is a non-universal anisotropy factor that characterizes the asymmetry between the intrinsic space and time directions of the statistical mechanical model with periodic boundary conditions. Using standard CFT results Affleck (1986); Blöte et al. (1986); Affleck (1988); Francesco et al. (2012), this implies that the bulk free energy density f(L)=F/(αtL)𝑓𝐿𝐹𝛼𝑡𝐿f(L)=F/(\alpha tL)italic_f ( italic_L ) = italic_F / ( italic_α italic_t italic_L ) – scales asZabalo et al. (2022)

f(L)=f(L=)πceff6L2+𝑓𝐿𝑓𝐿𝜋subscript𝑐eff6superscript𝐿2f(L)=f(L=\infty)-\frac{\pi c_{\rm{eff}}}{6L^{2}}+\dotsitalic_f ( italic_L ) = italic_f ( italic_L = ∞ ) - divide start_ARG italic_π italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT end_ARG start_ARG 6 italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + … (4)

for a cylindrical geometry of circumference L𝐿Litalic_L and circuit depth t𝑡titalic_t (when tLmuch-greater-than𝑡𝐿t\gg Litalic_t ≫ italic_L), where f(L=)𝑓𝐿f(L=\infty)italic_f ( italic_L = ∞ ) is the extensive bulk term, and the effective central charge

ceff=limk0dc(k)dk,subscript𝑐effsubscript𝑘0𝑑𝑐𝑘𝑑𝑘c_{\rm{eff}}=\lim_{k\to 0}\frac{{d}c(k)}{{d}k},italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = roman_lim start_POSTSUBSCRIPT italic_k → 0 end_POSTSUBSCRIPT divide start_ARG italic_d italic_c ( italic_k ) end_ARG start_ARG italic_d italic_k end_ARG , (5)

is a universal number that characterizes the log-CFT. Note that the actual central charge of the MIPT CFT is c=limk0c(k)=0𝑐subscript𝑘0𝑐𝑘0c=\lim_{k\to 0}c(k)=0italic_c = roman_lim start_POSTSUBSCRIPT italic_k → 0 end_POSTSUBSCRIPT italic_c ( italic_k ) = 0, since the partition function Z¯ksubscript¯𝑍𝑘\bar{Z}_{k}over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT becomes trivial Z¯k0=1subscript¯𝑍𝑘01\bar{Z}_{k\to 0}=1over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_k → 0 end_POSTSUBSCRIPT = 1 in the replica limit. Thus the free energy is trivial(=0)absent0(=0)( = 0 ) with no system size dependence and hence c=0𝑐0c=0italic_c = 0. In this bulk geometry, typical values of scaling dimensions are extracted from differences of higher (subleading) Lyapunov exponents, as shown in Ref. Zabalo et al., 2022.

II.2 Boundary CFT spectrum

In order to study boundary scaling properties, we now turn to a different geometry, consisting of an infinite strip of width L𝐿Litalic_L as shown in Fig 2. This coordinate system (x,t)superscript𝑥superscript𝑡(x^{\prime},t^{\prime})( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) with z=x+itsuperscript𝑧superscript𝑥𝑖superscript𝑡z^{\prime}=x^{\prime}+it^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_i italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT can be related to the upper half of the complex plane via the conformal transformation x+it=z=(iL/π)log(z)superscript𝑥𝑖superscript𝑡superscript𝑧𝑖𝐿𝜋𝑧x^{\prime}+it^{\prime}=z^{\prime}=(iL/\pi)\log(z)italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_i italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ( italic_i italic_L / italic_π ) roman_log ( italic_z ) with z=x+it𝑧𝑥𝑖𝑡z=x+ititalic_z = italic_x + italic_i italic_t. We introduce distinct boundary conditions on the bottom of the upper half plane, where the left boundary condition is depicted as red and labelled α𝛼\alphaitalic_α and the right boundary conditions is blue and labelled as β𝛽\betaitalic_β as shown in Fig. 2, corresponding to the insertion of BCC operator ΦαβsubscriptΦ𝛼𝛽\Phi_{\alpha\beta}roman_Φ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT at the origin. Upon applying the conformal transformation, this maps to distinct boundary conditions on the left and right edges of the strip (quantum circuit). In the strip geometry, the BCC operator ΦαβsubscriptΦ𝛼𝛽\Phi_{\alpha\beta}roman_Φ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT is inserted at imaginary time τ=𝜏\tau=-\inftyitalic_τ = - ∞, and changes the boundary conditions from α𝛼\alphaitalic_α to β𝛽\betaitalic_β from the left to the right side of the strip. As a result of the different (conformally invariant111Because the bulk is critical, any boundary condition implemented at the microscopic lattice scale of the circuit will at long scales result in a scale-invariant and conformally invariant boundary condition.) boundary conditions α𝛼\alphaitalic_α and β𝛽\betaitalic_β, the scaling of the averaged free energy density is given by Blöte et al. (1986); Cardy (1984b)

f(α|β)=f(L=)+fs(α|β)L+πhα|βL2πceff24L2.subscript𝑓conditional𝛼𝛽𝑓𝐿superscriptsubscript𝑓𝑠conditional𝛼𝛽𝐿𝜋subscriptconditional𝛼𝛽superscript𝐿2𝜋subscript𝑐eff24superscript𝐿2\displaystyle f_{(\alpha|\beta)}=f(L=\infty)+\frac{f_{{s}}^{(\alpha|\beta)}}{L% }+\frac{\pi h_{\alpha|\beta}}{L^{2}}-\frac{\pi c_{\rm{eff}}}{24L^{2}}.italic_f start_POSTSUBSCRIPT ( italic_α | italic_β ) end_POSTSUBSCRIPT = italic_f ( italic_L = ∞ ) + divide start_ARG italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_α | italic_β ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_L end_ARG + divide start_ARG italic_π italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG italic_π italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT end_ARG start_ARG 24 italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (6)

Compared to (4), there is an extra non-universal 1/L1𝐿1/L1 / italic_L dependence due to the presence of the surface free energy fs(α|β)superscriptsubscript𝑓𝑠conditional𝛼𝛽f_{{s}}^{(\alpha|\beta)}italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_α | italic_β ) end_POSTSUPERSCRIPT, and a boundary specific universal contribution from the scaling dimension hα|βsubscriptconditional𝛼𝛽h_{\alpha|\beta}italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT (=0absent0=0= 0, when α=β𝛼𝛽\alpha=\betaitalic_α = italic_β). As illustrated in Fig 2, this scaling dimension corresponds to the insertion of a BCC operator ΦαβsubscriptΦ𝛼𝛽\Phi_{\alpha\beta}roman_Φ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT at imaginary time τ=𝜏\tau=-\inftyitalic_τ = - ∞ which changes boundary conditions from α𝛼\alphaitalic_α to β𝛽\betaitalic_β from the left to the right side of the strip. Just like the effective central charge, the scaling dimension hα|βsubscriptconditional𝛼𝛽h_{\alpha|\beta}italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT is obtained as a derivative hα|β=limk0dhα|β(k)dksubscriptconditional𝛼𝛽subscript𝑘0𝑑subscriptconditional𝛼𝛽𝑘𝑑𝑘h_{\alpha|\beta}=\lim_{k\to 0}\frac{dh_{\alpha|\beta}(k)}{dk}italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT = roman_lim start_POSTSUBSCRIPT italic_k → 0 end_POSTSUBSCRIPT divide start_ARG italic_d italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT ( italic_k ) end_ARG start_ARG italic_d italic_k end_ARG from the replicated theory (and represents a typical scaling dimension). The surface free energy (fsα|β)superscriptsubscript𝑓𝑠conditional𝛼𝛽(f_{s}^{\alpha|\beta})( italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α | italic_β end_POSTSUPERSCRIPT ) contribution occurs in any statistical mechanics model with specified boundary ends, including for an example an Ising model on a strip.

Refer to caption
Figure 2: Conformal mapping: The conformal transformation (z=iLπlog(z)superscript𝑧𝑖𝐿𝜋logzz^{\prime}=\frac{iL}{\pi}\rm{log}(z)italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = divide start_ARG italic_i italic_L end_ARG start_ARG italic_π end_ARG roman_log ( roman_z )) of upper-half plane maps to infinite strip of width L𝐿Litalic_L. The two boundary ends α𝛼\alphaitalic_α (shown in red) and β𝛽\betaitalic_β (shown in blue) of infinite strip geometry results in the insertion of boundary observable ΦαβsubscriptΦ𝛼𝛽\Phi_{\alpha\beta}roman_Φ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT at the origin of the upper-half plane that separates the boundaries α𝛼\alphaitalic_α and β𝛽\betaitalic_β in the surface geometry. A semicircle (ACB) gets mapped to an equal-time line segment joining the boundary strip (ABsuperscriptAsuperscriptB\rm{A}^{\prime}\rm{B}^{\prime}roman_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT) as shown by the black dashed line. The respective boundary condition shown in Fig 1 will be denoted by (α,β)=(f,f),(f,a),𝛼𝛽𝑓𝑓𝑓𝑎(\alpha,\beta)=(f,f),(f,a),( italic_α , italic_β ) = ( italic_f , italic_f ) , ( italic_f , italic_a ) , and (a,b)𝑎𝑏(a,b)( italic_a , italic_b ), respectively – using notations consistent with Ref. Li et al. (2021b).

Using eq. (6), we see that we can extract numerically the BCC scaling dimensions hα|βsubscriptconditional𝛼𝛽h_{\alpha|\beta}italic_h start_POSTSUBSCRIPT italic_α | italic_β end_POSTSUBSCRIPT (belonging to the boundary spectrum of the CFT) from finite size scaling using various sets of boundary conditions (α,β)𝛼𝛽(\alpha,\beta)( italic_α , italic_β ). The boundary conditions that we will consider are guided by the underlying replica statistical mechanics model Jian et al. (2020). We will consider both “free” boundary conditions (corresponding to open boundary conditions and denoted by α=f𝛼𝑓\alpha=fitalic_α = italic_f), and different “fixed” boundary conditions, corresponding to fixing the boundary spins of the statistical mechanics model at the boundary. The degrees of freedom (“spins”) of the replicated statistical mechanics model are permutations of the replicas gSk𝑔subscript𝑆𝑘g\in S_{k}italic_g ∈ italic_S start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, and we will only consider two particular permutations corresponding to dissipation (identity permutation, label α=a𝛼𝑎\alpha=aitalic_α = italic_a), and entanglement (cyclic/swap permutation, label β=b𝛽𝑏\beta=bitalic_β = italic_b). This terminology is consistent with Ref. Li et al., 2021b.

II.3 Dissipative boundary setup (α,β)=(f,a)𝛼𝛽𝑓𝑎(\alpha,\beta)=(f,a)( italic_α , italic_β ) = ( italic_f , italic_a )

We first introduce dissipators at a boundary end to implement the boundary condition shown in Fig 1(b). A single qubit maximally mixed depolarizing channel or dephasing along the z𝑧zitalic_z-axis are used to model dissipation. For random Haar circuits, dephasing and depolarizing channels are expected to flow – in the renormalization group (RG) sense – to the same conformally invariant boundary condition. These are added after each time period as shown in Fig 1(b). The resulting dynamics is described by the quantum channel

𝒩t(𝒟)(ρ)=𝒟(𝒯𝐭𝒟(𝒯𝟐(𝒟(𝒯𝟏ρ𝒯𝟏))𝒯𝟐)𝒯𝐭),superscriptsubscript𝒩𝑡𝒟𝜌𝒟subscript𝒯𝐭𝒟subscript𝒯2𝒟subscript𝒯1𝜌superscriptsubscript𝒯1superscriptsubscript𝒯2superscriptsubscript𝒯𝐭\mathcal{N}_{t}^{(\mathcal{D})}(\rho)=\mathcal{D}(\mathcal{T}_{\bf{t}}\dots% \mathcal{D}(\mathcal{T}_{\bf{2}}(\mathcal{D}(\mathcal{T}_{\bf{1}}\rho\mathcal{% T}_{\bf{1}}^{\dagger}))\mathcal{T}_{\bf{2}}^{\dagger})\dots\mathcal{T}_{\bf{t}% }^{\dagger}),caligraphic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( caligraphic_D ) end_POSTSUPERSCRIPT ( italic_ρ ) = caligraphic_D ( caligraphic_T start_POSTSUBSCRIPT bold_t end_POSTSUBSCRIPT … caligraphic_D ( caligraphic_T start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ( caligraphic_D ( caligraphic_T start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT italic_ρ caligraphic_T start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) ) caligraphic_T start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) … caligraphic_T start_POSTSUBSCRIPT bold_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) , (7)

where 𝒟𝒟\mathcal{D}caligraphic_D is the single qubit depolarizing/dephasing mapNielsen and Chuang (2010) that acts on the boundary qubit x=L𝑥𝐿x=Litalic_x = italic_L. The maximally mixed depolarizing channel maps ρ𝒟(ρ)=I/2𝜌𝒟𝜌𝐼2\rho\to\mathcal{D}(\rho)=I/2italic_ρ → caligraphic_D ( italic_ρ ) = italic_I / 2, which models absolute decoherence and the dephasing channel maps ρ𝒟(ρ)=PρP+PρP𝜌𝒟𝜌subscript𝑃𝜌subscript𝑃subscript𝑃𝜌subscript𝑃\rho\to\mathcal{D}(\rho)=P_{\uparrow}\rho P_{\uparrow}+P_{\downarrow}\rho P_{\downarrow}italic_ρ → caligraphic_D ( italic_ρ ) = italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_ρ italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_P start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT italic_ρ italic_P start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT, which amounts to adding measurements on this qubit and discarding the measurement outcomes Weinstein et al. (2022); Li and Claassen (2023). This random circuit geometry will be denoted by the boundary conditions (α,β)=(f,a)𝛼𝛽𝑓𝑎(\alpha,\beta)=(f,a)( italic_α , italic_β ) = ( italic_f , italic_a ), and in the statistical mechanics model language it corresponds to fixing boundary spins at the right boundary to the identity permutation.

Mapping this boundary condition back to the half-plane, we see that the (f,a)𝑓𝑎(f,a)( italic_f , italic_a ) boundary condition corresponds to a setup in which one measures the spins on the left half-line at the final time, and computes the resulting Born probability, Tr(|ψ𝐦~ψ|𝐦~)Trsubscriptket𝜓~𝐦subscriptbra𝜓~𝐦\mathrm{Tr}(|\psi\rangle_{\mathbf{\tilde{m}}}\langle\psi|_{\mathbf{\tilde{m}}})roman_Tr ( | italic_ψ ⟩ start_POSTSUBSCRIPT over~ start_ARG bold_m end_ARG end_POSTSUBSCRIPT ⟨ italic_ψ | start_POSTSUBSCRIPT over~ start_ARG bold_m end_ARG end_POSTSUBSCRIPT ), where 𝐦~~𝐦\mathbf{\tilde{m}}over~ start_ARG bold_m end_ARG is a bit-string of measurement outcomes for both the mid-circuit measurements in the spacetime bulk of the circuit and the measurements on the left half-system on the final time-step. This quantity receives a nonuniversal surface contribution (which would exist even if all the sites were independent) and a subleading part capturing the long-range correlations between the outcomes of distant measurements.

Another observable that contains direct information about the (typical) scaling dimension hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT turns out to be the Shannon entropy of the measurement record of the circuit at early time, i.e. that of a shallow circuit. To see this consider a “sideways variant” of Fig. 2, where the roles of space and time are exchanged: Specifically, it is convenient to first consider a circuit with periodic boundary conditions in space of “large” circumference L𝐿Litalic_L, and “small” depth t𝑡titalic_t, so that tLmuch-less-than𝑡𝐿t\ll Litalic_t ≪ italic_L. Then choose as initial condition the state labeled by the boundary condition f𝑓fitalic_f (which can be represented, e.g., by a simple product state); the state of the circuit appearing at depth t𝑡titalic_t, the ‘final’ time, which contains the physical qubits, is represented by boundary condition a𝑎aitalic_a 222For both of these interpretations of the boundary conditions f𝑓fitalic_f and a𝑎aitalic_a, as (simple product) initial state and final state containing the physical qubits, respectively, see Fig. 2 (a), and the corresponding text of Ref. Li et al., 2021b.. Now consider the partition function Z𝐦subscript𝑍𝐦Z_{\mathbf{m}}italic_Z start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT of the so-defined shallow circuit which equals333paragraph below Eq. 2 the Born probability for the measurement record, p𝐦=subscript𝑝𝐦absentp_{\mathbf{m}}=italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = Z𝐦subscript𝑍𝐦Z_{\mathbf{m}}italic_Z start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT. For this shallow circuit we are considering now, where tLmuch-less-than𝑡𝐿t\ll Litalic_t ≪ italic_L, the corresponding Born-probability-averaged free energy density, which by definition, Eq. 3, is the corresponding shallow-circuit Shannon entropy density, has the same form444 see, e.g., Ref. Li et al., 2021b as Eq. 6, except that the roles of space and time are exchanged,

limLlnZ𝐦¯αtL=limL1αtL𝐦p𝐦lnp𝐦=subscript𝐿¯subscript𝑍𝐦𝛼𝑡𝐿subscript𝐿1𝛼𝑡𝐿subscript𝐦subscript𝑝𝐦subscript𝑝𝐦absent\displaystyle-\lim_{L\to\infty}{\overline{\ln Z_{\mathbf{m}}}\over\alpha tL}=-% \lim_{L\to\infty}{1\over\alpha tL}\sum_{\mathbf{m}}\ p_{\mathbf{m}}\ln p_{% \mathbf{m}}=- roman_lim start_POSTSUBSCRIPT italic_L → ∞ end_POSTSUBSCRIPT divide start_ARG over¯ start_ARG roman_ln italic_Z start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT end_ARG end_ARG start_ARG italic_α italic_t italic_L end_ARG = - roman_lim start_POSTSUBSCRIPT italic_L → ∞ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_α italic_t italic_L end_ARG ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT roman_ln italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT =
=f(t=)+fs(f|a)t+πhf|at2πceff24t2,absent𝑓𝑡superscriptsubscript𝑓𝑠conditional𝑓𝑎𝑡𝜋subscriptconditional𝑓𝑎superscript𝑡2𝜋subscript𝑐eff24superscript𝑡2\displaystyle=f(t=\infty)+{f_{s}^{(f|a)}\over t}+{\pi\ h_{f|a}\over t^{2}}-{% \pi\ c_{\rm eff}\over 24t^{2}},\qquad= italic_f ( italic_t = ∞ ) + divide start_ARG italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_f | italic_a ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_t end_ARG + divide start_ARG italic_π italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT end_ARG start_ARG italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG italic_π italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT end_ARG start_ARG 24 italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (8)

where O¯𝐦=𝐦p𝐦O𝐦subscript¯𝑂𝐦subscript𝐦subscript𝑝𝐦subscript𝑂𝐦\overline{O}_{\bf m}=\sum_{\bf m}p_{\bf m}O_{\bf m}over¯ start_ARG italic_O end_ARG start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT refers to average over quantum trajectories, and f(t=)𝑓𝑡f(t=\infty)italic_f ( italic_t = ∞ ) and fs(f|a)superscriptsubscript𝑓𝑠conditional𝑓𝑎f_{s}^{(f|a)}italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_f | italic_a ) end_POSTSUPERSCRIPT are non-universal. Eq. 8 shows that the 1t21superscript𝑡2{1\over t^{2}}divide start_ARG 1 end_ARG start_ARG italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG-decay of the Shannon entropy density at early times is directly affected by the exponent hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT, a universal signature of the sensitivity of this entropy to the initial state f𝑓fitalic_f. At long times, the circuit “forgets” about the initial state f𝑓fitalic_f, and at finite spatial circumference L𝐿Litalic_L the behavior in Eq. 4 obtains in the opposite limit of a deep circuit where tLmuch-greater-than𝑡𝐿t\gg Litalic_t ≫ italic_L, in the steady state, for the case of periodic spatial boundary conditions we are currently considering. (Open spatial boundary conditions do not modify the early-time tLmuch-less-than𝑡𝐿t\ll Litalic_t ≪ italic_L behaviorLi et al. (2021b), Eq. 8, but the late-time form Eq. 6 obtains in the opposite limit of a deep circuit, tLmuch-greater-than𝑡𝐿t\gg Litalic_t ≫ italic_L, in the steady state.)

II.4 Boundary ancillas measurement-induced entanglement setup (α,β)=(a,b)𝛼𝛽𝑎𝑏(\alpha,\beta)=(a,b)( italic_α , italic_β ) = ( italic_a , italic_b )

In order to implement different boundary conditions, we introduce ancilla qubits at left and right edges Li et al. (2021b) as shown in Fig. 1(c). Every time step, we introduce two fresh system qubits that individually form a Bell pair with an ancilla qubit and inject them into the system at the first (x=1𝑥1x=1italic_x = 1) and last (x=L𝑥𝐿x=Litalic_x = italic_L) qubit location. Then after the evolution of system qubits for one time period we take out the first and last qubit to store them as ancillas with no further action on them. After that we repeat by introducing two fresh pre-entangled qubits at the boundary ends again, and so on. Thus for a circuit of depth t𝑡titalic_t, we introduce 2t2𝑡2t2 italic_t ancilla qubits. Tracing over the ancillas at a boundary effectively implements dissipation, corresponding to α=a𝛼𝑎\alpha=aitalic_α = italic_a.

We now compute the entanglement between the right and left ancillas, while measuring all physical qubits – corresponding to a free f𝑓fitalic_f boundary condition on the top layer of the circuit555see Ref. Li et al., 2021b, Appendix C, and Ref. Popp et al., 2005.. This measurement-induced entanglement (MIE) Popp et al. (2005); Lin et al. (2023) between right and left boundaries effectively implements a change in boundary conditions: one of the boundaries is traced over (boundary condition α=a𝛼𝑎\alpha=aitalic_α = italic_a), while the other is subject to a partial trace (corresponding to a cyclic permutation of replicas, boundary condition β=b𝛽𝑏\beta=bitalic_β = italic_b). In the statistical mechanics language, this forces the insertion of a domain wall propagating vertically between the two ends of the strip. The MIE is then directly given by the free energy cost of changing this boundary condition:

SMIEαLt=f(a|b)f(f|f)=πha|bL2+subscript𝑆MIE𝛼𝐿𝑡subscript𝑓conditional𝑎𝑏subscript𝑓conditional𝑓𝑓𝜋subscriptconditional𝑎𝑏superscript𝐿2\displaystyle\frac{S_{\rm{MIE}}}{\alpha Lt}=f_{(a|b)}-f_{(f|f)}=\frac{\pi h_{a% |b}}{L^{2}}+\dotsdivide start_ARG italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT end_ARG start_ARG italic_α italic_L italic_t end_ARG = italic_f start_POSTSUBSCRIPT ( italic_a | italic_b ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = divide start_ARG italic_π italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + … (9)

A “sideway” version of this geometry, where the space and the time directions of the circuit were exchanged, was considered in Ref. Li et al., 2021b. We thus see that this geometry allows us to extract the “entanglement” scaling dimension ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT which controls the scaling of the entanglement entropy at criticality. In particular, in previous works ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT was typically extracted directly from the logarithmic scaling of entanglement entropy (of the physical qubits) at criticality SA2ha|blnLAsimilar-tosubscript𝑆𝐴2subscriptconditional𝑎𝑏subscript𝐿𝐴S_{A}\sim 2h_{a|b}\ln L_{A}italic_S start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∼ 2 italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT roman_ln italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT, which exhibits the coefficient γ𝛾\gammaitalic_γ of the logarithm in Eq. (1), γ=2ha|b𝛾2subscriptconditional𝑎𝑏\gamma=2h_{a|b}italic_γ = 2 italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT. As already mentioned, we can unfortunately not use this boundary condition with Haar random circuits as the Hilbert space dimension grows with time as one adds ancilla qubits as depicted in Fig. 1.

II.5 Numerical analysis

In the rest of this paper, we apply this boundary transfer matrix approach numerically. We average free energies over 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT sampled quantum trajectories where each trajectory is evolved up to time t=10L𝑡10𝐿t=10Litalic_t = 10 italic_L, after an initial equilibration time (4Lsimilar-toabsent4𝐿\sim 4L∼ 4 italic_L). The sampling complexity of quantum trajectories limits us to inspect small system sizes, even with state-of-the-art computing platforms. However we still manage to do larger systems, both for Clifford and Haar, as compared to the previous work with periodic boundaryZabalo et al. (2022). All results are shown at criticality p=pc𝑝subscript𝑝𝑐p=p_{c}italic_p = italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, with pcMOIM=0.5superscriptsubscript𝑝𝑐MOIM0.5p_{c}^{\rm{MOIM}}=0.5italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_MOIM end_POSTSUPERSCRIPT = 0.5, pcDC=0.205superscriptsubscript𝑝𝑐DC0.205p_{c}^{\rm{DC}}=0.205italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_DC end_POSTSUPERSCRIPT = 0.205, pcDH=0.14superscriptsubscript𝑝𝑐DH0.14p_{c}^{\rm{DH}}=0.14italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_DH end_POSTSUPERSCRIPT = 0.14, and pcC=0.1596superscriptsubscript𝑝𝑐𝐶0.1596p_{c}^{C}=0.1596italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT = 0.1596, for MOIM, dual Clifford, dual Haar, and Clifford circuits, respectively Gullans and Huse (2020b); Zabalo et al. (2020). Note that adding dissipators at boundary does not influence saturation time of free energy density. However it results in additional Monte Carlo sampling of trajectories for Haar circuit as described in Appendix D ( Fig 12). This further limits the system sizes and thus results in larger error bars for the universal and non-universal quantities as compared to the Clifford circuit. This also leads to larger error bars for the dissipator boundary as compared to the open boundary condition, even within the Haar circuit. The space-time asymmetries for different monitored circuits are characterized by the anisotropy parameter (α𝛼\alphaitalic_α) and we extract this from the ratio of space and time correlators Zabalo et al. (2022) which gives αMOIM=αDC=αDH=1subscript𝛼MOIMsubscript𝛼DCsubscript𝛼DH1\alpha_{\rm{MOIM}}=\alpha_{\rm{DC}}=\alpha_{\rm{DH}}=1italic_α start_POSTSUBSCRIPT roman_MOIM end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT roman_DC end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT roman_DH end_POSTSUBSCRIPT = 1 for MOIM, dual Clifford, dual Haar circuits (which are all isotropic), and αC=0.61subscript𝛼𝐶0.61\alpha_{C}=0.61italic_α start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = 0.61 for the Clifford random circuit. All error bars are estimated using a bootstrap analysis where the data is bootstrapped over 1000 samples. To improve the estimate of universal quantities we use standard double fitting procedure where we successively remove small system sizes (L<Lmin𝐿subscript𝐿minL<L_{\rm{min}}italic_L < italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT) from the fit which in turn accounts for the leading order correction to the averaged free energy density.

III Measurement-only Ising model

To demonstrate the validity and accuracy of the boundary transfer matrix spectrum approach to MIPTs, we first consider the measurement-only Ising model (MOIM) Nahum and Skinner (2020); Lang and Büchler (2020) where the conformal spectrum can be derived analytically.

III.1 Measurement-only dynamics

Measurement-only circuits are comprised of different non-commuting measurement operators and are free from random unitary gates Nahum and Skinner (2020); Lang and Büchler (2020); Ippoliti et al. (2021); Lavasani et al. (2021b); Sang and Hsieh (2021); Sang et al. (2021); Sriram et al. (2022). We consider the measurement-only Ising model (MOIM) which describes measurement-only dynamics using non-commuting competitive measurement operators, σixsubscriptsuperscript𝜎𝑥𝑖\sigma^{x}_{i}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and σizσi+1zsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1\sigma^{z}_{i}\sigma^{z}_{i+1}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT, acting on sites i{1,2,,L}𝑖12𝐿i\in\{1,2,\dots,L\}italic_i ∈ { 1 , 2 , … , italic_L } of a one-dimensional spin-1/2 chain. The dynamics involve the projective measurement operator M[O]𝑀delimited-[]𝑂M[O]italic_M [ italic_O ] that measures the observable O𝑂Oitalic_O, which then maps the state |ψket𝜓|\psi\rangle| italic_ψ ⟩ onto the eigenspace of O𝑂Oitalic_O, i.e., M[O]|ψ=Pλ|ψψ|Pλ|ψ𝑀delimited-[]𝑂ket𝜓subscript𝑃𝜆ket𝜓quantum-operator-product𝜓subscript𝑃𝜆𝜓M[O]|\psi\rangle=\frac{P_{\lambda}|\psi\rangle}{\sqrt{\langle\psi|P_{\lambda}|% \psi\rangle}}italic_M [ italic_O ] | italic_ψ ⟩ = divide start_ARG italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT | italic_ψ ⟩ end_ARG start_ARG square-root start_ARG ⟨ italic_ψ | italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT | italic_ψ ⟩ end_ARG end_ARG, with probability pλ=ψ|Pλ|ψsubscript𝑝𝜆quantum-operator-product𝜓subscript𝑃𝜆𝜓p_{\lambda}=\langle\psi|P_{\lambda}|\psi\rangleitalic_p start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = ⟨ italic_ψ | italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT | italic_ψ ⟩ where λ𝜆\lambdaitalic_λ is the eigenvalue of the corresponding eigenspace of O𝑂Oitalic_O and Pλsubscript𝑃𝜆P_{\lambda}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is the projection operator onto the eigenspace of O𝑂Oitalic_O with eigenvalue λ𝜆\lambdaitalic_λ. For MOIM, the observable O𝑂Oitalic_O is either σixsubscriptsuperscript𝜎𝑥𝑖\sigma^{x}_{i}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT or σizσi+1zsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1\sigma^{z}_{i}\sigma^{z}_{i+1}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT and the corresponding projection operator Pλsubscript𝑃𝜆P_{\lambda}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is either I+λσix2𝐼𝜆subscriptsuperscript𝜎𝑥𝑖2\frac{I+\lambda\sigma^{x}_{i}}{2}divide start_ARG italic_I + italic_λ italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG or I+λσizσi+1z2𝐼𝜆subscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖12\frac{I+\lambda\sigma^{z}_{i}\sigma^{z}_{i+1}}{2}divide start_ARG italic_I + italic_λ italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG where λ𝜆\lambdaitalic_λ takes value ±1plus-or-minus1\pm 1± 1. In each discrete time step, the measurement operators M[σizσi+1z]𝑀delimited-[]subscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1M[\sigma^{z}_{i}\sigma^{z}_{i+1}]italic_M [ italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ] (Mezzabsentsuperscriptsubscript𝑀𝑒𝑧𝑧\equiv M_{e}^{zz}≡ italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z italic_z end_POSTSUPERSCRIPT) and M[σix](Mix)annotated𝑀delimited-[]subscriptsuperscript𝜎𝑥𝑖absentsuperscriptsubscript𝑀𝑖𝑥M[\sigma^{x}_{i}](\equiv M_{i}^{x})italic_M [ italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] ( ≡ italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ) acts randomly on each edge e=(i,i+1)𝑒𝑖𝑖1e=(i,i+1)italic_e = ( italic_i , italic_i + 1 ) and site i𝑖iitalic_i with a probability 1p1𝑝1-p1 - italic_p and p𝑝pitalic_p, respectively. The layer of Mezzsuperscriptsubscript𝑀𝑒𝑧𝑧M_{e}^{zz}italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z italic_z end_POSTSUPERSCRIPT measurements precedes all probable Mixsuperscriptsubscript𝑀𝑖𝑥M_{i}^{x}italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT operators. The initial state is set to |ψ0=|+++|\psi_{0}\rangle=|++\dots+\rangle| italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ = | + + ⋯ + ⟩ where |+=|0+|12ketket0ket12|+\rangle=\frac{|0\rangle+|1\rangle}{\sqrt{2}}| + ⟩ = divide start_ARG | 0 ⟩ + | 1 ⟩ end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG is the eigenstate of σxsuperscript𝜎𝑥\sigma^{x}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT operator and thus the resulting dynamics leaves the state invariant under the operator 𝒞=i=1Lσix𝒞superscriptsubscriptproduct𝑖1𝐿superscriptsubscript𝜎𝑖𝑥\mathcal{C}=\prod_{i=1}^{L}\sigma_{i}^{x}caligraphic_C = ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT which describes the symmetry of this model. The transfer matrix evolving the system for a single time step then acts as

|ψ(t+1)=M𝐢xM𝐞zz|ψ(t),ket𝜓𝑡1tensor-productsuperscriptsubscript𝑀𝐢𝑥subscriptsuperscript𝑀𝑧𝑧𝐞ket𝜓𝑡\displaystyle|\psi(t+1)\rangle=M_{\bf{i}}^{x}\otimes M^{zz}_{\bf{e}}|\psi(t)\rangle,| italic_ψ ( italic_t + 1 ) ⟩ = italic_M start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ⊗ italic_M start_POSTSUPERSCRIPT italic_z italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_e end_POSTSUBSCRIPT | italic_ψ ( italic_t ) ⟩ , (10)

where 𝐞𝐞\bf{e}bold_e and 𝐢𝐢\bf{i}bold_i are the sets of all edges and sites, respectively, on which the measurement operations (Mezzsuperscriptsubscript𝑀𝑒𝑧𝑧M_{e}^{zz}italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z italic_z end_POSTSUPERSCRIPT and Mixsuperscriptsubscript𝑀𝑖𝑥M_{i}^{x}italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT) act. It has been shown that the model exhibits an entanglement phase transition between two area law phases and the physics at the critical point is described by bond percolation (which is a CFT with central charge c=0𝑐0c=0italic_c = 0, as it should to describe a MIPT).

III.2 Percolation mapping and replicated statistical mechanics model

Refer to caption
Figure 3: Percolation mapping of the MOIM: (a) Each σizσi+1zsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1\sigma^{z}_{i}\sigma^{z}_{i+1}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT measurement (blue edge) along an edge maps to a horizontal bond and the absence of a σixsubscriptsuperscript𝜎𝑥𝑖\sigma^{x}_{i}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (green circles) measurement corresponds to a vertical bond. This defines a percolation cluster (shown by dark black line). A given time step comprises of a layer of σizσi+1zsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1\sigma^{z}_{i}\sigma^{z}_{i+1}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT and then σixsubscriptsuperscript𝜎𝑥𝑖\sigma^{x}_{i}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT measurement operations with probability (1p)1𝑝(1-p)( 1 - italic_p ) and p𝑝pitalic_p respectively, that evolves the initial state (X1,X2,X3,X4,X5)subscript𝑋1subscript𝑋2subscript𝑋3subscript𝑋4subscript𝑋5(X_{1},X_{2},X_{3},X_{4},X_{5})( italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ) from t=0𝑡0t=0italic_t = 0 to time t=5,whereXσxformulae-sequence𝑡5whereXsuperscript𝜎xt=5\rm{,~{}where~{}}X\equiv\sigma^{x}italic_t = 5 , roman_where roman_X ≡ italic_σ start_POSTSUPERSCRIPT roman_x end_POSTSUPERSCRIPT. At the last time step, one artificially measures all sites using σxsuperscript𝜎𝑥\sigma^{x}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT measurement operators. The resulting configuration has parameter Ns=30subscript𝑁𝑠30N_{s}=30italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 30 (number of space-time sites), Nvb=11subscript𝑁vb11N_{\rm{vb}}=11italic_N start_POSTSUBSCRIPT roman_vb end_POSTSUBSCRIPT = 11 (number of vertical bonds), and Nc=11subscript𝑁𝑐11N_{c}=11italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 11 (number of percolation clusters, including single-site clusters denoted by isolated black dots). (b) In this example we fix the order of measurement from left to right end in each time step. The total number of random measurements Nrand=16subscript𝑁rand16N_{\rm{rand}}=16italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT = 16 (marked by stars) is independent of the choice of order in which the measurements are performed. All measured links (vertical or horizontal) which are not marked by a star in this panel correspond to deterministic measurements (meaning the Born probability of that particular measurement outcome is 1111). This example satisfies the relation between percolation clusters and number of random measurements stated in Eq. 11.

We now review the mapping of this model onto bond percolation Nahum and Skinner (2020); Lang and Büchler (2020), and generalize it to fully characterize the associated replicated statistical mechanics model Zk=𝐦p𝐦k+1subscript𝑍𝑘subscript𝐦superscriptsubscript𝑝𝐦𝑘1Z_{k}=\sum_{{\bf m}}p_{\bf m}^{k+1}italic_Z start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k + 1 end_POSTSUPERSCRIPT. First, note that to each realization of the circuit, one can associate a bond percolation configuration as shown in Fig. 3(a), where a measurement along an edge Mezzsuperscriptsubscript𝑀𝑒𝑧𝑧M_{e}^{zz}italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z italic_z end_POSTSUPERSCRIPT corresponds to a horizontal bond along that edge, whereas the absence of a local measurement Mixsuperscriptsubscript𝑀𝑖𝑥M_{i}^{x}italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT on a site maps to a vertical bond. This provides a one-to-one correspondence between percolation configurations and measurement locations. As we now show, the free energy is entirely given by the properties of this percolation configuration.

First, note that the MOIM is a stabilizer (Clifford) circuit, so that each measurement outcome is either fully deterministic (determined outcomes), or fully random (with equally probable outcomes). For example let us suppose we have a stabilizer state |+,+ket|+,+\rangle| + , + ⟩. Then measuring σ1zσ2zsubscriptsuperscript𝜎𝑧1subscriptsuperscript𝜎𝑧2\sigma^{z}_{1}\sigma^{z}_{2}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT will result in either (|00+|11)/2ket00ket112(|00\rangle+|11\rangle)/\sqrt{2}( | 00 ⟩ + | 11 ⟩ ) / square-root start_ARG 2 end_ARG or (|01+|10)/2ket01ket102(|01\rangle+|10\rangle)/\sqrt{2}( | 01 ⟩ + | 10 ⟩ ) / square-root start_ARG 2 end_ARG, each with probability 1/2121/21 / 2. However if we measure σ1xsubscriptsuperscript𝜎𝑥1\sigma^{x}_{1}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for the same state, we get back |+,+ket|+,+\rangle| + , + ⟩ with probability 1. We call the former as random and the latter to be a deterministic measurement outcome. Note that for MOIM whether an outcome is random or deterministic will depend on the order in which measurements are performed in a given layer of time but the total number of random measurements is independent of the choice of order. We choose the order of measurements from left to right end in each time step. As a result, the Born probability of a given trajectory is given by p𝐦=(12)Nrandsubscript𝑝𝐦superscript12subscript𝑁randp_{\bf m}=\left(\frac{1}{2}\right)^{N_{\rm rand}}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, where Nrandsubscript𝑁randN_{\rm rand}italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT is the total number of random (non-deterministic) measurements. This means that the free energy is given by F=N¯randln(2)𝐹subscript¯𝑁rand2F=\overline{N}_{\rm rand}\ln(2)italic_F = over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT roman_ln ( 2 ), where N¯randsubscript¯𝑁rand\overline{N}_{\rm rand}over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT is the average number of random measurements in those circuits.

Refer to caption
Figure 4: Boundary spectrum of MOIM: We implement the boundary condition in Fig 1 to obtain the respective boundary spectrum of MOIM. (a) The plot of f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1/L1𝐿1/L1 / italic_L clearly respect the scaling form stated in Eq. 6 (for Lmin{10,12,14,16,18}subscript𝐿min1012141618L_{\rm{min}}\in\{10,12,14,16,18\}italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ∈ { 10 , 12 , 14 , 16 , 18 }) as shown by the dashed blue line. In the inset, we perform the double fitting procedure to extract effective central charge ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and free energy surface term f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT, by successively removing the smallest system size and then fitting the data with LLmin𝐿subscript𝐿minL\geq L_{\rm{min}}italic_L ≥ italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT. This gives us ceff=1.0(1)subscript𝑐eff1.01c_{\rm{eff}}=1.0(1)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 1.0 ( 1 ) and f(f|f)s=0.4671(7)superscriptsubscript𝑓conditional𝑓𝑓𝑠0.46717f_{(f|f)}^{s}=-0.4671(7)italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT = - 0.4671 ( 7 ) as shown by the red dashed line. (b) The plot of (f(f|a)f(f|f))Lsubscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓𝐿(f_{(f|a)}-f_{(f|f)})L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ) italic_L scales linearly with 1/L1𝐿1/L1 / italic_L as shown by the dashed blue line, and from fitting the values of slope and intercept, we obtain the dissipator scaling dimension hf|a=0.048(6)subscriptconditional𝑓𝑎0.0486h_{f|a}=0.048(6)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.048 ( 6 ) and fδs=0.27subscript𝑓𝛿𝑠0.27f_{\delta s}=0.27italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.27, as shown with their respective red dashed lines in the inset plots. (c) The plot of SMIEsubscript𝑆MIES_{\textrm{MIE}}italic_S start_POSTSUBSCRIPT MIE end_POSTSUBSCRIPT vs 1/L1𝐿1/L1 / italic_L clearly follows the trend stated in Eq. 9 and we find the entanglement scaling dimension ha|b=0.097(2)subscriptconditional𝑎𝑏0.0972h_{a|b}=0.097(2)italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = 0.097 ( 2 ) using the red dashed line in the inset plot. The fit of SMIEsubscript𝑆MIES_{\textrm{MIE}}italic_S start_POSTSUBSCRIPT MIE end_POSTSUBSCRIPT include a leading correction term coefficient A1=0.37subscript𝐴10.37A_{1}=0.37italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.37. All the data shown above for the MOIM have pc=0.5subscript𝑝𝑐0.5p_{c}=0.5italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.5, α=1𝛼1\alpha=1italic_α = 1, and L{10,12,14,16,18,20,24,28}𝐿1012141618202428L\in\{10,12,14,16,18,20,24,28\}italic_L ∈ { 10 , 12 , 14 , 16 , 18 , 20 , 24 , 28 }. We sample =5×106absent5superscript106=5\times 10^{6}= 5 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT trajectories to find the average value of free energy density.

Next, we notice that the number of random measurements is purely determined in terms of the geometrical properties of the percolation clusters. Namely, we find by inspection that

Nrand=2(NsNvbNc),subscript𝑁rand2subscript𝑁𝑠subscript𝑁vbsubscript𝑁𝑐\displaystyle N_{\rm{rand}}=2(N_{s}-N_{\rm{vb}}-N_{c}),italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT = 2 ( italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_N start_POSTSUBSCRIPT roman_vb end_POSTSUBSCRIPT - italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) , (11)

where Nssubscript𝑁𝑠N_{s}italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, Nvbsubscript𝑁vbN_{\rm{vb}}italic_N start_POSTSUBSCRIPT roman_vb end_POSTSUBSCRIPT, and Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are the total number of sites, vertical bonds, and disconnected clusters in the percolation configuration. An example illustrating the validity of this expression is provided in Fig. 3 where we fix the order of measurements from left to right end in each time step. For this particular choice, the random and deterministic measurement locations are shown in Fig. 3(b). So for a fixed measurement trajectory with all measurement outcomes set to 1111, we write the evolution of stabilizer generators for one time step from t=2𝑡2t=2italic_t = 2 to t=3𝑡3t=3italic_t = 3 in Fig. 3(b), following the convention Xσx𝑋superscript𝜎𝑥X\equiv\sigma^{x}italic_X ≡ italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT and Zσz𝑍superscript𝜎𝑧Z\equiv\sigma^{z}italic_Z ≡ italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT to denote these generators. The stabilizer generators at t=2𝑡2t=2italic_t = 2 are given by {X1,X2,X3,X4,X5}subscript𝑋1subscript𝑋2subscript𝑋3subscript𝑋4subscript𝑋5\{X_{1},X_{2},X_{3},X_{4},X_{5}\}{ italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT }; this set evolves to {X1X2,Z1Z2,X3X4,Z3Z4,X5}subscript𝑋1subscript𝑋2subscript𝑍1subscript𝑍2subscript𝑋3subscript𝑋4subscript𝑍3subscript𝑍4subscript𝑋5\{X_{1}X_{2},Z_{1}Z_{2},X_{3}X_{4},Z_{3}Z_{4},X_{5}\}{ italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_Z start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT } upon measuring σ1zσ2zsubscriptsuperscript𝜎𝑧1subscriptsuperscript𝜎𝑧2\sigma^{z}_{1}\sigma^{z}_{2}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and σ3zσ4zsubscriptsuperscript𝜎𝑧3subscriptsuperscript𝜎𝑧4\sigma^{z}_{3}\sigma^{z}_{4}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT. Clearly both measurements are random in accordance with the two site example discussed earlier. Next we measure σ2xsubscriptsuperscript𝜎𝑥2\sigma^{x}_{2}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and σ4xsubscriptsuperscript𝜎𝑥4\sigma^{x}_{4}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT, which results in the stabilizer generators {X1,X2,X3X4,Z3Z4,X5}subscript𝑋1subscript𝑋2subscript𝑋3subscript𝑋4subscript𝑍3subscript𝑍4subscript𝑋5\{X_{1},X_{2},X_{3}X_{4},Z_{3}Z_{4},X_{5}\}{ italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_Z start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT }, at the end of time t=3𝑡3t=3italic_t = 3. The measurement on site 2222 is random while the measurement on last qubit is deterministic since the qubit already is in the |+ket|+\rangle| + ⟩ state. Doing this for all time steps in turn leads to the total random measurements that obey the formula in Eq. 11. Although we do not have a formal proof of this formula, we have checked its validity numerically for very large circuits. This equation can further be broken down into purely extensive contributions, and terms including non-extensive universal corrections. The first two terms scale extensively (O(Lt)𝑂𝐿𝑡O(Lt)italic_O ( italic_L italic_t )) and will therefore contribute to the bulk free energy only. Dropping these non-universal extensive contributions, we thus find that

F2N¯cln2,similar-to𝐹2subscript¯𝑁𝑐2F\sim-2\overline{N}_{c}\ln 2,italic_F ∼ - 2 over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_ln 2 , (12)

where N¯csubscript¯𝑁𝑐\overline{N}_{c}over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the average number of percolation clusters at criticality. This quantity can be computed using the standard Fortuin-Kasteleyn (FK) Newman (1994) mapping between percolation and the Q1𝑄1Q\to 1italic_Q → 1 limit of the Q𝑄Qitalic_Q-state Potts model. Using these standard percolation results, we find that N¯c=ddQFPotts|Q=1subscript¯𝑁𝑐evaluated-at𝑑𝑑𝑄superscript𝐹Potts𝑄1\overline{N}_{c}=-\left.\frac{d}{dQ}F^{\rm{Potts}}\right|_{Q=1}over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = - divide start_ARG italic_d end_ARG start_ARG italic_d italic_Q end_ARG italic_F start_POSTSUPERSCRIPT roman_Potts end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT italic_Q = 1 end_POSTSUBSCRIPT where FPottssuperscript𝐹PottsF^{\rm{Potts}}italic_F start_POSTSUPERSCRIPT roman_Potts end_POSTSUPERSCRIPT is the free energy of the classical Q𝑄Qitalic_Q-state Potts model. This provides a direct relation between the free energy of the circuit (Shannon entropy of measurement record) and that of the classical Potts model. In fact, using the same reasoning as above, one can show that the replicated partition function Zk=𝐦p𝐦k+1subscript𝑍𝑘subscript𝐦superscriptsubscript𝑝𝐦𝑘1Z_{k}=\sum_{{\bf m}}p_{\bf m}^{k+1}italic_Z start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k + 1 end_POSTSUPERSCRIPT maps onto a Potts model with Q=4k𝑄superscript4𝑘Q=4^{k}italic_Q = 4 start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT (up to non-universal contributions), with Q1𝑄1Q\to 1italic_Q → 1 in the replica limit k0𝑘0k\to 0italic_k → 0. We can then use standard CFT results for the Potts model to infer universal quantities for the MOIM transition. For example, the central charge of the Potts model Francesco et al. (2012) is given by cPotts=16m(m+1)superscript𝑐Potts16𝑚𝑚1c^{\rm{Potts}}=1-\frac{6}{m(m+1)}italic_c start_POSTSUPERSCRIPT roman_Potts end_POSTSUPERSCRIPT = 1 - divide start_ARG 6 end_ARG start_ARG italic_m ( italic_m + 1 ) end_ARG where m=πarccos(Q2)1𝑚𝜋𝑄21m=\frac{\pi}{\arccos(\frac{\sqrt{Q}}{2})}-1italic_m = divide start_ARG italic_π end_ARG start_ARG roman_arccos ( divide start_ARG square-root start_ARG italic_Q end_ARG end_ARG start_ARG 2 end_ARG ) end_ARG - 1. We thus find that the effective central charge is given by

ceff=ln4ddQcPotts|Q=1=53ln22π0.96.subscript𝑐effevaluated-at4𝑑𝑑𝑄superscript𝑐Potts𝑄15322𝜋similar-to-or-equals0.96\left.c_{\rm{eff}}=\ln 4\frac{d}{dQ}c^{\rm{Potts}}\right|_{Q=1}=\frac{5\sqrt{3% }\ln 2}{2\pi}\simeq 0.96.italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = roman_ln 4 divide start_ARG italic_d end_ARG start_ARG italic_d italic_Q end_ARG italic_c start_POSTSUPERSCRIPT roman_Potts end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT italic_Q = 1 end_POSTSUBSCRIPT = divide start_ARG 5 square-root start_ARG 3 end_ARG roman_ln 2 end_ARG start_ARG 2 italic_π end_ARG ≃ 0.96 . (13)

Scaling dimensions can be identified through this mapping as well, and fit into the “Kac table” Francesco et al. (2012) hr,s=((r(m+1)sm)21)/(4m(m+1))subscript𝑟𝑠superscript𝑟𝑚1𝑠𝑚214𝑚𝑚1h_{r,s}=((r(m+1)-sm)^{2}-1)/(4m(m+1))italic_h start_POSTSUBSCRIPT italic_r , italic_s end_POSTSUBSCRIPT = ( ( italic_r ( italic_m + 1 ) - italic_s italic_m ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) / ( 4 italic_m ( italic_m + 1 ) ).

We now turn to a numerical analysis of the MOIM in the various geometries summarized in Fig. 1. We will first implement the open (free) boundary conditions, Fig 1(a)), to extract ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT. Then we introduce dissipation at one end of the boundary to evaluate the dissipator scaling dimension (hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT). Finally to extract the entanglement scaling dimension ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT, we implement the entangled system-ancilla setup (as described in fig 1(c)). We make use of stabilizer formalism Aaronson and Gottesman (2004); Gottesman (1998); Nahum et al. (2017); Li et al. (2019); Krastanov et al. (2022), which allows for efficient classical simulation of large quantum circuits.

III.3 Effective central charge (ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT)

We first compute the numerical value of the effective central charge in the cylindrical geometry setup with periodic boundary conditions. The plot of free energy density (f)𝑓(f)( italic_f ), shown in Appendix A (Fig. 8(a)), follows a straight line when plotted against 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as expected from Eq 4 and the slope gives ceff=0.96(1)subscript𝑐eff0.961c_{\rm{eff}}=0.96(1)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.96 ( 1 ). This is in good agreement with the analytic prediction in Eq (13). We also extracted a precise value of f(L=)=0.55718(2)𝑓𝐿0.557182f(L=\infty)=0.55718(2)italic_f ( italic_L = ∞ ) = 0.55718 ( 2 ), using the linear double fitting procedure. We now implement MOIM with free boundary ends as shown in Fig 1(a), corresponding to (α,β)=(f,f)𝛼𝛽𝑓𝑓(\alpha,\beta)=(f,f)( italic_α , italic_β ) = ( italic_f , italic_f ). The free energy density scales in accordance with Eq. 6 where hf|f=0subscriptconditional𝑓𝑓0h_{f|f}=0italic_h start_POSTSUBSCRIPT italic_f | italic_f end_POSTSUBSCRIPT = 0 (as expectedLi et al. (2021b)) and the resulting fit gives ceff=1.0(1)subscript𝑐eff1.01c_{\rm{eff}}=1.0(1)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 1.0 ( 1 ) as shown in Fig 4(a). Note that this estimate is less accurate than the periodic boundary conditions case, largely because of the non-universal surface term f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT which adds a fitting parameter. In order to reduce the number of fitting parameters, we use the extensive free energy term f(L=)𝑓𝐿f(L=\infty)italic_f ( italic_L = ∞ ) obtained from periodic boundary conditions in Eq. 6.

III.4 Dissipator scaling dimension (hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT)

Next, we implement boundary depolarizing/dephasing channels as shown in fig 1(b), to extract the dissipator scaling dimension hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT. In order to get rid of the extensive bulk contribution, we plot the free energy difference

f(f|a)f(f|f)=fδsL+πhf|aL2,subscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓subscript𝑓𝛿𝑠𝐿𝜋subscriptconditional𝑓𝑎superscript𝐿2f_{(f|a)}-f_{(f|f)}=\frac{f_{\delta s}}{L}+\frac{\pi h_{f|a}}{L^{2}},italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = divide start_ARG italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_L end_ARG + divide start_ARG italic_π italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (14)

where fδs=f(f|a)sf(f|f)ssubscript𝑓𝛿𝑠subscriptsuperscript𝑓𝑠conditional𝑓𝑎subscriptsuperscript𝑓𝑠conditional𝑓𝑓f_{\delta s}=f^{s}_{(f|a)}-f^{s}_{(f|f)}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT. The plot of (f(f|a)f(f|f))L(f_{(f|a)}-f_{(f|f}))L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f end_POSTSUBSCRIPT ) ) italic_L vs 1L1𝐿\frac{1}{L}divide start_ARG 1 end_ARG start_ARG italic_L end_ARG displays a linear trend with 1/L1𝐿1/L1 / italic_L as shown in 4(b). Now using the double fitting procedure we find the boundary scaling dimension hf|a(L)=0.0480.432L2subscriptconditional𝑓𝑎𝐿0.0480.432superscript𝐿2h_{f|a}(L)=0.048-\frac{0.432}{L^{2}}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_L ) = 0.048 - divide start_ARG 0.432 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and the non-universal free energy surface term fδs=0.27+0.06L2subscript𝑓𝛿𝑠0.270.06superscript𝐿2f_{\delta s}=0.27+\frac{0.06}{L^{2}}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.27 + divide start_ARG 0.06 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. We conclude that hf|a=0.048(6)subscriptconditional𝑓𝑎0.0486h_{f|a}=0.048(6)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.048 ( 6 ) and fδs=0.275(1)subscript𝑓𝛿𝑠0.2751f_{\delta s}=0.275(1)italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.275 ( 1 ). Note that the addition of dissipators still preserves the global symmetry 𝒞𝒞\mathcal{C}caligraphic_C. We also observe that the scaling dimension remains unchanged with the type of dissipators since dephasing along zlimit-from𝑧z-italic_z -axis is equivalent to depolarizing in this model.

The mapping onto bond percolation can be generalized in the presence of boundary dissipators. We find that boundary dissipation induces additional random (non-deterministic) measurements that are associated with the hulls of percolation clusters touching the boundary, see Appendix B. The BCC operator associated with changing the fugacity of such boundary hulls was identified in Ref. Jacobsen and Saleur (2008). Combining these results, we find that hf|a=38πln20.048subscriptconditional𝑓𝑎38𝜋2similar-to-or-equals0.048h_{f|a}=\frac{\sqrt{3}}{8\pi}\ln 2\simeq 0.048italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 8 italic_π end_ARG roman_ln 2 ≃ 0.048 (see Appendix B), in very good agreement with our numerical results.

Refer to caption
Figure 5: Boundary spectrum of dual Clifford circuit:We implement the boundary condition in Fig 1 to obtain the respective boundary spectrum of dual Clifford circuit. (a) The plot of f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1/L1𝐿1/L1 / italic_L clearly respect the scaling form stated in Eq. 6 (for Lmin{10,12,14,16,18}subscript𝐿min1012141618L_{\rm{min}}\in\{10,12,14,16,18\}italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ∈ { 10 , 12 , 14 , 16 , 18 }) as shown by the dashed blue line. In the inset, we perform the double fitting procedure to extract effective central charge ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and free energy surface term f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT, by successively removing the smallest system size and then fitting the data with LLmin𝐿subscript𝐿minL\geq L_{\rm{min}}italic_L ≥ italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT. This gives us ceff=0.32(3)subscript𝑐eff0.323c_{\rm{eff}}=0.32(3)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.32 ( 3 ) and f(f|f)s=0.0615(2)superscriptsubscript𝑓conditional𝑓𝑓𝑠0.06152f_{(f|f)}^{s}=-0.0615(2)italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT = - 0.0615 ( 2 ) as shown by the red dashed line. (b) The plot of (f(f|a)f(f|f))Lsubscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓𝐿(f_{(f|a)}-f_{(f|f)})L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ) italic_L scales linearly with 1/L1𝐿1/L1 / italic_L as shown by the dashed blue line, and from fitting the value of slope and intercept we obtain the dissipator scaling dimension hf|a=0.029(2)subscriptconditional𝑓𝑎0.0292h_{f|a}=0.029(2)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.029 ( 2 ) and fδs=0.0479subscript𝑓𝛿𝑠0.0479f_{\delta s}=0.0479italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.0479 as shown with their respective red dashed lines in the inset plots. (c) The plot of SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT vs 1/L1𝐿1/L1 / italic_L clearly follows the trend stated in Eq. 9 and we find the entanglement scaling dimension ha|b=0.519(8)subscriptconditional𝑎𝑏0.5198h_{a|b}=0.519(8)italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = 0.519 ( 8 ) using the red dashed line in the inset plot. The fit of SMIEsubscript𝑆MIES_{\textrm{MIE}}italic_S start_POSTSUBSCRIPT MIE end_POSTSUBSCRIPT include a leading correction term coefficient A1=1.833subscript𝐴11.833A_{1}=-1.833italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 1.833. All the data shown above for the dual Clifford circuit have pc=0.205subscript𝑝𝑐0.205p_{c}=0.205italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.205, α=1𝛼1\alpha=1italic_α = 1, and L{10,12,14,16,18,20,24,28}𝐿1012141618202428L\in\{10,12,14,16,18,20,24,28\}italic_L ∈ { 10 , 12 , 14 , 16 , 18 , 20 , 24 , 28 } except for SMIEsubscript𝑆MIES_{\textrm{MIE}}italic_S start_POSTSUBSCRIPT MIE end_POSTSUBSCRIPT where L=28𝐿28L=28italic_L = 28 is absent. We sample =5×106absent5superscript106=5\times 10^{6}= 5 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT trajectories to find the average value of free energy density.

III.5 Entanglement scaling dimension (ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT)

Finally, the boundary scaling dimension ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT has been determined both numerically and analytically for the MOIM with cylindrical geometryLang and Büchler (2020), with ha|b=ln4ddQh1,2Potts|Q=10.096subscriptconditional𝑎𝑏evaluated-at4𝑑𝑑𝑄subscriptsuperscriptPotts12𝑄1similar-to-or-equals0.096h_{a|b}=\left.\ln 4\frac{d}{dQ}h^{\rm{Potts}}_{1,2}\right|_{Q=1}\simeq 0.096italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = roman_ln 4 divide start_ARG italic_d end_ARG start_ARG italic_d italic_Q end_ARG italic_h start_POSTSUPERSCRIPT roman_Potts end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT | start_POSTSUBSCRIPT italic_Q = 1 end_POSTSUBSCRIPT ≃ 0.096. This critical exponent corresponds to counting clusters crossing the strip. Here, we extract it using the boundary ancilla setup of Fig. 1(c). We compute the measurement-induced entanglement (SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT) between ancilla qubits present on the left and right ends of the boundary where all system qubits after the last time step are measured. The result in Fig. 4(c) agrees with the scaling form Eq. 9 and we extract the entanglement scaling dimension ha|b(L)=0.0970.06L2subscriptconditional𝑎𝑏𝐿0.0970.06superscript𝐿2h_{a|b}(L)=0.097-\frac{0.06}{L^{2}}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT ( italic_L ) = 0.097 - divide start_ARG 0.06 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. We conclude that ha|b=0.097(2)subscriptconditional𝑎𝑏0.0972h_{a|b}=0.097(2)italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = 0.097 ( 2 ), which is in good agreement with the theory prediction. Hence these results completely support our approach to obtain the boundary conformal spectrum. We thus can now generalize it to the more general unitary-measurement dynamics.

IV Dual Clifford Random Circuit

The dual Clifford random circuit consists of two-qubit dual-unitary Clifford gates and local measurement operators that project along the computational basis states.

The two-site dual-unitary Clifford gates comprise those Clifford gates which obey the duality relation which makes them unitary both in space and time directions Bertini et al. (2019). The dual-unitary condition is satisfied by the SWAP and the iSWAP classes of the two qubit Clifford operators. This contains a total of 5760 (576 SWAP + 5184 iSWAP) gate operations Barends et al. (2014). Using these gates fixes the anisotropy parameter α=1𝛼1\alpha=1italic_α = 1, allowing us to extract critical properties more accurately. We then follow the boundary transfer matrix approach outlined in Sec. II.

IV.1 Effective central charge (ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT)

We will first start by discussing the calculation of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in cylindrical geometry (periodic boundary conditions). In Appendix A (Fig 8(b)), we find that the free energy density f𝑓fitalic_f shows a clear linear dependence when plotted against 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in accordance with the CFT result stated in Eq. 4. This yields ceff=0.349(3)subscript𝑐eff0.3493c_{\rm{eff}}=0.349(3)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.349 ( 3 ) and f(L=)=0.131574(6)𝑓𝐿0.1315746f(L=\infty)=0.131574(6)italic_f ( italic_L = ∞ ) = 0.131574 ( 6 ). We now implement dual Clifford circuit with free boundary ends where we find ceff=0.32(3)subscript𝑐eff0.323c_{\rm{eff}}=0.32(3)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.32 ( 3 ) from the plot between f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1/L1𝐿1/L1 / italic_L as shown in fig 5(a). More precisely, using a double fitting procedure, we find ceff(L)=0.328.74L2subscript𝑐eff𝐿0.328.74superscript𝐿2c_{\rm{eff}}(L)=0.32-\frac{8.74}{L^{2}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_L ) = 0.32 - divide start_ARG 8.74 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and f(f|f)s(L)=0.06150.0569L2superscriptsubscript𝑓conditional𝑓𝑓𝑠𝐿0.06150.0569superscript𝐿2f_{(f|f)}^{s}(L)=-0.0615-\frac{0.0569}{L^{2}}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( italic_L ) = - 0.0615 - divide start_ARG 0.0569 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. The plot respects the scaling form given in Eq. 6 and the value of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT is in agreement with the one extracted from periodic boundary conditions. As in the MOIM case, ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT has relatively large error bars as a consequence of the free energy surface term. Last, we find that f(f|f)s=0.0615(2)superscriptsubscript𝑓conditional𝑓𝑓𝑠0.06152f_{(f|f)}^{s}=-0.0615(2)italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT = - 0.0615 ( 2 ).

IV.2 Dissipator scaling dimension (hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT)

We now implement boundary dissipation as shown in Fig 1(b), to extract the dissipator scaling dimension hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT. The plot of (f(f|a)f(f|f))L(f_{(f|a)}-f_{(f|f}))L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f end_POSTSUBSCRIPT ) ) italic_L vs 1L1𝐿\frac{1}{L}divide start_ARG 1 end_ARG start_ARG italic_L end_ARG displays a linear dependence against 1/L1𝐿1/L1 / italic_L as shown in Fig 5(b). Now using the double fitting procedure we find the boundary scaling dimension hf|a(L)=0.0290.430L2subscriptconditional𝑓𝑎𝐿0.0290.430superscript𝐿2h_{f|a}(L)=0.029-\frac{0.430}{L^{2}}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_L ) = 0.029 - divide start_ARG 0.430 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and the non-universal free energy surface term fδs=0.0479+0.0666L2subscript𝑓𝛿𝑠0.04790.0666superscript𝐿2f_{\delta s}=0.0479+\frac{0.0666}{L^{2}}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.0479 + divide start_ARG 0.0666 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. We conclude that hf|a=0.029(2)subscriptconditional𝑓𝑎0.0292h_{f|a}=0.029(2)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.029 ( 2 ) and fδs=0.0479(3)subscript𝑓𝛿𝑠0.04793f_{\delta s}=0.0479(3)italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.0479 ( 3 ). Note that the scaling dimension remains invariant with the type of dissipator used at the boundary, consistent with the fact that both depolarizing and dephasing channels correspond to the same conformally invariant boundary condition α=a𝛼𝑎\alpha=aitalic_α = italic_a.

IV.3 Entanglement scaling dimension (ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT)

The entanglement scaling dimension ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT is known for Clifford circuit with cylindrical geometryLi et al. (2018, 2019); Skinner et al. (2019); Li et al. (2021b) from the coefficient of the subsytem entanglement entropy.Here, we extract it from the boundary ancilla setup shown in Fig 1(c).

We compute the measurement-induced entanglement SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT between ancilla qubits present on the left and right ends of the boundary where all system qubits after the last time step are measured. The plot in Fig 5(c) clearly respects the scaling form 9 and we extract the entanglement scaling dimension ha|b(L)=0.519+0.116L2subscriptconditional𝑎𝑏𝐿0.5190.116superscript𝐿2h_{a|b}(L)=0.519+\frac{0.116}{L^{2}}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT ( italic_L ) = 0.519 + divide start_ARG 0.116 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. The entanglement scaling dimension ha|b=0.519(8)subscriptconditional𝑎𝑏0.5198h_{a|b}=0.519(8)italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = 0.519 ( 8 ) agrees with existing results, and provides yet another check of our approach.

While we focused on dual-unitary Clifford circuits in this section, we also checked that we obtain consistent (but less accurate) results for regular Clifford circuits (see Appendix C). This confirms the expectation that Clifford and dual-unitary Clifford monitored circuits are in the same universality class.

V Dual Haar Random Circuit

We now come to the case of Haar random circuits undergoing random and local projective measurements. The qubit chain evolves in time under a bricklayer sequence of discrete timesteps involving two-qubit entangling gates Ui,i+1subscript𝑈𝑖𝑖1U_{i,i+1}italic_U start_POSTSUBSCRIPT italic_i , italic_i + 1 end_POSTSUBSCRIPT and local σizsubscriptsuperscript𝜎𝑧𝑖\sigma^{z}_{i}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT measurements with probability p𝑝pitalic_p as depicted in Fig. 1. As in the previous section, to avoid the error associated with computing the space time anisotropy, we consider dual-unitary circuits by choosing only the gates Ui,i+1subscript𝑈𝑖𝑖1U_{i,i+1}italic_U start_POSTSUBSCRIPT italic_i , italic_i + 1 end_POSTSUBSCRIPT that are unitary in space and time such that α=1𝛼1\alpha=1italic_α = 1. The dual unitary gates are given by,

U=eiϕ(u+u)V[J](vv+),𝑈superscript𝑒𝑖italic-ϕtensor-productsubscript𝑢subscript𝑢𝑉delimited-[]𝐽tensor-productsubscript𝑣subscript𝑣U=e^{i\phi}(u_{+}\otimes u_{-})\cdot V[J]\cdot(v_{-}\otimes v_{+}),italic_U = italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT ( italic_u start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ⊗ italic_u start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ) ⋅ italic_V [ italic_J ] ⋅ ( italic_v start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ⊗ italic_v start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) , (15)

where ϕ,Jitalic-ϕ𝐽\phi,J\in\mathbb{R}italic_ϕ , italic_J ∈ blackboard_R are chosen randomly from [0,π)0𝜋[0,\pi)[ 0 , italic_π ) and u±,v±SU(2)subscript𝑢plus-or-minussubscript𝑣plus-or-minusSU2u_{\pm},v_{\pm}\in\rm{SU(2)}italic_u start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ∈ roman_SU ( 2 ) are randomly chosen using the Haar measure, and

V[J]=exp[iπ4(σxσx+σyσy+Jσzσz)].𝑉delimited-[]𝐽𝑖𝜋4tensor-productsuperscript𝜎𝑥superscript𝜎𝑥tensor-productsuperscript𝜎𝑦superscript𝜎𝑦tensor-product𝐽superscript𝜎𝑧superscript𝜎𝑧V[J]=\exp\left[-i\frac{\pi}{4}\left(\sigma^{x}\otimes\sigma^{x}+\sigma^{y}% \otimes\sigma^{y}+J\sigma^{z}\otimes\sigma^{z}\right)\right].italic_V [ italic_J ] = roman_exp [ - italic_i divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ( italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ⊗ italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT ⊗ italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT + italic_J italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ⊗ italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ) ] . (16)

The statistical self-duality of these models under rotations forces α=1𝛼1\alpha=1italic_α = 1, allowing for more accurate estimate of their critical properties.

Refer to caption
Figure 6: Haar dual unitaries free energy with open boundaries. The free energy density f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT with open boundary condition vs L𝐿Litalic_L. The data fits well with the expected scaling form given in Eq.(6) with open boundaries f(f|f)(L)=f(L=)+0.0251/L0.09417/L2subscript𝑓conditional𝑓𝑓𝐿𝑓𝐿0.0251𝐿0.09417superscript𝐿2f_{(f|f)}(L)=f(L=\infty)+0.0251/L-0.09417/L^{2}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ( italic_L ) = italic_f ( italic_L = ∞ ) + 0.0251 / italic_L - 0.09417 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT allowing us to extract ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and f(f|f)ssubscriptsuperscript𝑓𝑠conditional𝑓𝑓f^{s}_{(f|f)}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT. The top and bottom insets show the double fitting procedure we used to extract ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and f(f|f)ssubscriptsuperscript𝑓𝑠conditional𝑓𝑓f^{s}_{(f|f)}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT by successively removing the smallest system size(Lminsubscript𝐿minL_{\rm{min}}italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT) from the fit. For f(f|f)ssubscriptsuperscript𝑓𝑠conditional𝑓𝑓f^{s}_{(f|f)}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT, we find f(f|f)s(Lmin)=0.09352+0.0431/Lmin2subscriptsuperscript𝑓𝑠conditional𝑓𝑓subscript𝐿min0.093520.0431superscriptsubscript𝐿min2f^{s}_{(f|f)}(L_{\rm{min}})=0.09352+0.0431/L_{\rm{min}}^{2}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ( italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) = 0.09352 + 0.0431 / italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT shown by the dotted line in the top inset. For ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, we find ceff(Lmin)=0.274.92/Lmin2subscript𝑐effsubscript𝐿min0.274.92subscriptsuperscript𝐿2minc_{\rm{eff}}(L_{\rm{min}})=0.27-4.92/L^{2}_{\rm{min}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) = 0.27 - 4.92 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT shown be the dotted line in the bottom inset. We conclude ceff=0.27(2)subscript𝑐eff0.272c_{\rm{eff}}=0.27(2)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.27 ( 2 ) and f(f|f)s=0.09352(2)subscriptsuperscript𝑓𝑠conditional𝑓𝑓0.093522f^{s}_{(f|f)}=0.09352(2)italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = 0.09352 ( 2 ). We used 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT trajectories for averaging each data point in this plot.

V.1 Effective central charge (ceff)subscript𝑐eff(c_{\rm{eff}})( italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT )

First, we consider the dynamics of the model with open boundary conditions shown in Fig. 1 (a). As discussed in Sec. II, the effective central charge can be obtained from the free energy density f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT expected to obey the scaling form given in Eq. (6). We note that numerically probing the CFT with OBC is costly as it requires averaging the quantities over a large number of trajectories compared to that with cylindrical geometry. For the results presented in this section, we obtained 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT samples for statistical averaging, similar to the stabilizer simulation presented before. In Fig. 6, we show f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT at late times (tLmuch-greater-than𝑡𝐿t\gg Litalic_t ≫ italic_L) vs L𝐿Litalic_L for L=8𝐿8L=8italic_L = 8 to 20202020. In addition to the leading 1/L1𝐿1/L1 / italic_L scaling, f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT exhibits the expected 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT correction. As suggested in Eq.(6), the coefficient of 1/L1𝐿1/L1 / italic_L term gives f(f|f)ssubscriptsuperscript𝑓𝑠conditional𝑓𝑓f^{s}_{(f|f)}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT, while that of the 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is related to ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT. To obtain estimates of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and f(f|f)ssubscriptsuperscript𝑓𝑠conditional𝑓𝑓f^{s}_{(f|f)}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT, we use the double fitting procedure explained previously to fit with the forms, ceff(L)=ceff(L=)+b/L2subscript𝑐eff𝐿subscript𝑐eff𝐿𝑏superscript𝐿2c_{\rm{eff}}(L)=c_{\rm{eff}}(L=\infty)+b/L^{2}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_L ) = italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_L = ∞ ) + italic_b / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and f(f|f)s(L)=f(f|f)s(L=)+b~/L2subscriptsuperscript𝑓𝑠conditional𝑓𝑓𝐿subscriptsuperscript𝑓𝑠conditional𝑓𝑓𝐿~𝑏superscript𝐿2f^{s}_{(f|f)}(L)=f^{s}_{(f|f)}(L=\infty)+\tilde{b}/L^{2}italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ( italic_L ) = italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ( italic_L = ∞ ) + over~ start_ARG italic_b end_ARG / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT shown in bottom and top insets of Fig. 6 respectively. This double fitting procedure gives an estimate of ceff=0.27(2)subscript𝑐eff0.272c_{\rm{eff}}=0.27(2)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.27 ( 2 ) which matches with the previous result for periodic boundary condition (ceff=0.24(2)subscript𝑐eff0.242c_{\rm{eff}}=0.24(2)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.24 ( 2 )) obtained in Ref. Zabalo et al., 2022. We also estimate the surface free energy, f(f|f)s=0.09352(2)subscriptsuperscript𝑓𝑠conditional𝑓𝑓0.093522f^{s}_{(f|f)}=0.09352(2)italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = 0.09352 ( 2 ). We note that we used the extensive bulk parameter f(L=)=0.8902(2)𝑓𝐿0.89022f(L=\infty)=0.8902(2)italic_f ( italic_L = ∞ ) = 0.8902 ( 2 ), which is obtained using periodic boundary conditions as we did for the other circuit models.

Refer to caption
Figure 7: Haar dissipator free energy and its difference with open boundaries. The dissipator free energies fϕ(L)subscript𝑓italic-ϕ𝐿f_{\phi}(L)italic_f start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ( italic_L ) displays expected dependence on L𝐿Litalic_L from Eq. (6). For system sizes up to L=14𝐿14L=14italic_L = 14 we ran 6×1086superscript1086\times 10^{8}6 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT trajectories, including Monte Carlo samples. For L=16𝐿16L=16italic_L = 16, we ran 6×1076superscript1076\times 10^{7}6 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT trajectories. The difference in free energy densities displays the correct dependence as expected from Eq. (14), ff|a(L)ff|f(L)=0.004L+0.077L2subscript𝑓conditional𝑓𝑎𝐿subscript𝑓conditional𝑓𝑓𝐿0.004𝐿0.077superscript𝐿2f_{f|a}(L)-f_{f|f}(L)=\frac{0.004}{L}+\frac{0.077}{L^{2}}italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_L ) - italic_f start_POSTSUBSCRIPT italic_f | italic_f end_POSTSUBSCRIPT ( italic_L ) = divide start_ARG 0.004 end_ARG start_ARG italic_L end_ARG + divide start_ARG 0.077 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG allowing us to extract fssubscript𝑓𝑠f_{s}italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT (Inset top) and hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT (inset bottom). We find that fδs(Lmin)=0.003+0.566Lmin2subscript𝑓𝛿𝑠subscript𝐿min0.0030.566superscriptsubscript𝐿min2f_{\delta s}(L_{\rm min})=-0.003+\frac{0.566}{L_{\rm min}^{2}}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT ( italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) = - 0.003 + divide start_ARG 0.566 end_ARG start_ARG italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and hf|a(Lmin)=0.052.26Lmin2subscriptconditional𝑓𝑎subscript𝐿min0.052.26superscriptsubscript𝐿min2h_{f|a}(L_{\rm min})=0.05-\frac{2.26}{L_{\rm min}^{2}}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) = 0.05 - divide start_ARG 2.26 end_ARG start_ARG italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.
Table 1: The non-universal and universal parameters for various circuit types at critical probability pcsubscript𝑝𝑐p_{c}italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the non-universal parameter extracted from the scaling form of free energy density and measurement-induced entanglement are bulk free energy f(L=)𝑓𝐿f(L=\infty)italic_f ( italic_L = ∞ ), open surface free energy f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT, and depolarizer surface free energy f(f|a)ssuperscriptsubscript𝑓conditional𝑓𝑎𝑠f_{(f|a)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT. The universal parameters are effective central charge ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT with periodic boundary condition (PBC) and open boundary condition (OBC), boundary scaling dimensions; hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT and ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT.
Non-universal parameters Universal parameters
Circuit type

pcsubscript𝑝𝑐p_{c}italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT

f(L=)𝑓𝐿f(L=\infty)italic_f ( italic_L = ∞ )

f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT

f(f|a)ssuperscriptsubscript𝑓conditional𝑓𝑎𝑠f_{(f|a)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT

ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT (PBC)

ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT (OBC)

hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT

ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT

MOIM

0.5

0.55718(2)

-0.4671(7)

-0.1919(7)

0.96(1)

1.0(1)

0.048(6)

0.097(2)

Dual Clifford

0.205

0.131574(6)

-0.0615(2)

-0.0136(3)

0.349(3)

0.32(3)

0.029(2)

0.519(8)

Clifford

0.1596

0.170469(6)

-0.0749(3)

-0.0202(3)

0.375(5)

0.32(4)

0.034(2)

0.534(2)

Dual Haar

0.14

0.8902(2)

0.09352(2)

0.094(8)

0.24(2)

0.27(2)

0.05(1)

V.2 Dissipator scaling dimension (hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT)

We next apply a “dissipator” to one end of the open boundary conditions. In particular in the odd time steps, the qubit at the right end of the chain (i.e. site x=L𝑥𝐿x=Litalic_x = italic_L) is subjected to the dephasing channel along the z-axis which can be rewritten as,

𝒟(ρ)=PρP+PρP=pdρ+(1pd)σzρσz,𝒟𝜌subscript𝑃𝜌subscript𝑃subscript𝑃𝜌subscript𝑃subscript𝑝𝑑𝜌1subscript𝑝𝑑subscript𝜎𝑧𝜌subscript𝜎𝑧\mathcal{D}(\rho)=P_{\uparrow}\rho P_{\uparrow}+P_{\downarrow}\rho P_{% \downarrow}=p_{d}\rho+(1-p_{d})\sigma_{z}\rho\sigma_{z},caligraphic_D ( italic_ρ ) = italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_ρ italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_P start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT italic_ρ italic_P start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_ρ + ( 1 - italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_ρ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (17)

where pd=12subscript𝑝𝑑12p_{d}=\frac{1}{2}italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG. The monitored dynamics with the dissipator can be described by a stochastic master equation, which describes the evolution of the density matrix, ρ˙=ρ˙𝜌𝜌\dot{\rho}=\mathcal{L}\rhoover˙ start_ARG italic_ρ end_ARG = caligraphic_L italic_ρ where the Liouville superoperator \mathcal{L}caligraphic_L under the Lindblad approximation takes the form,

ρ˙=ρ=c^ρc^12{c^c^,ρ^}.˙𝜌𝜌^𝑐𝜌superscript^𝑐12^𝑐superscript^𝑐^𝜌\dot{\rho}=\mathcal{L}\rho=\hat{c}\rho\hat{c}^{\dagger}-\frac{1}{2}\{\hat{c}% \hat{c}^{\dagger},\hat{\rho}\}.over˙ start_ARG italic_ρ end_ARG = caligraphic_L italic_ρ = over^ start_ARG italic_c end_ARG italic_ρ over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG { over^ start_ARG italic_c end_ARG over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , over^ start_ARG italic_ρ end_ARG } . (18)

Here the Lindblad operator c^=σz1/(2dt)^𝑐subscript𝜎𝑧12𝑑𝑡\hat{c}=\sigma_{z}\sqrt{1/(2dt)}over^ start_ARG italic_c end_ARG = italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT square-root start_ARG 1 / ( 2 italic_d italic_t ) end_ARG.

Writing the dissipator in this form allows us to Monte Carlo sample many quantum trajectories from a single dephasing channel on the end qubit, and a fixed set of gates and measurement locations on the other qubits. Using this method we can study the dissipative dynamics without having to resort to simulating the full density matrix which is much more numerically challenging(e.g. ρ𝜌\rhoitalic_ρ requires 22Lsuperscript22𝐿2^{2L}2 start_POSTSUPERSCRIPT 2 italic_L end_POSTSUPERSCRIPT numbers, whereas storing the statevector only requires 2Lsuperscript2𝐿2^{L}2 start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT. Nonetheless, due to the large number of circuit realizations required to handle the open boundary conditions, how this Monte Carlo sampling is done on top of that, which requires a Monte Carlo average for each circuit sample, remains a non-trivial task.

As the only quantity we are aiming to compute is the free energy density, we can utilize its dependence on the Monte Carlo sampling “time” (denoted by τ𝜏\tauitalic_τ and we stress its not a real physical time) ff|a(t,L;τ)subscript𝑓conditional𝑓𝑎𝑡𝐿𝜏f_{f|a}(t,L;\tau)italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_t , italic_L ; italic_τ ) to extrapolate the Monte Carlo averages to τ𝜏\tau\rightarrow\inftyitalic_τ → ∞. We find that ff|a(t,L;τ)subscript𝑓conditional𝑓𝑎𝑡𝐿𝜏f_{f|a}(t,L;\tau)italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_t , italic_L ; italic_τ ) converges like 1/τ1𝜏1/\tau1 / italic_τ at sufficiently late τ𝜏\tauitalic_τ. Instead of converging each circuit sample in Monte Carlo time, we instead work at a fixed Monte Carlo realization of the dissipator, and then average over samples of the circuit. This produces a smooth function that we extrapolate to τ𝜏\tau\rightarrow\inftyitalic_τ → ∞ using a fourth order polynomial in 1/τ1𝜏1/\tau1 / italic_τ. Hence, using this τ𝜏\tauitalic_τ dependence, we extract the Monte Carlo average using a reasonable number (600similar-toabsent600\sim 600∼ 600) of Monte Carlo samples. This procedure is demonstrated in more details in Appendix D.

With the dissipative free energy in hand f(f|a)subscript𝑓conditional𝑓𝑎f_{(f|a)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT, we compute its difference with free boundaries f(f|a)f(f|f)=fδsL+πhf|aL2subscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓subscript𝑓𝛿𝑠𝐿𝜋subscriptconditional𝑓𝑎superscript𝐿2f_{(f|a)}-f_{(f|f)}=\frac{f_{\delta s}}{L}+\frac{\pi h_{f|a}}{L^{2}}italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = divide start_ARG italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_L end_ARG + divide start_ARG italic_π italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG where fδs=f(f|a)sf(f|f)ssubscript𝑓𝛿𝑠subscriptsuperscript𝑓𝑠conditional𝑓𝑎subscriptsuperscript𝑓𝑠conditional𝑓𝑓f_{\delta s}=f^{s}_{(f|a)}-f^{s}_{(f|f)}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT. The results and scaling of the data is shown in Fig. 7. After performing a similar double fitting procedure, we obtain f(f|a)s=0.094(8)subscriptsuperscript𝑓𝑠conditional𝑓𝑎0.0948f^{s}_{(f|a)}=0.094(8)italic_f start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT = 0.094 ( 8 ) and find hf|a=0.05(1)subscriptconditional𝑓𝑎0.051h_{f|a}=0.05(1)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.05 ( 1 ). This appears to be different from the dissipator scaling dimension measured in the Clifford case, as expected. We note that this critical exponent is new and was not measured before, and it would be interesting to find other geometries and quantities to extract it. Of course, the boundary-ancilla setup used to measure the entanglement scaling dimension ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT in the Clifford case cannot be implemented in the Haar case as the number of ancillas grows linearly with time, so our transfer matrix approach cannot be used in that case.

VI Conclusions

In this work, we investigated the boundary conformal properties of MIPTs for various circuit models as summarized in Table 1 by introducing a numerical boundary transfer matrix approach. This in turn characterizes the CFT that describes these transitions. It further solidifies the fact that Haar and Clifford circuit do not belong to the same universality class. The extracted effective central charges in the cylindrical and infinite strip geometry agree with each other, validating the overall transfer matrix approach as an efficient way to probe MIPTs numerically. This is further validated by the analytically tractable case of the MOIM. In this work, we extracted the scaling dimension associated with two specific BCCBCC{\rm BCC}roman_BCC operators, namely the dissipator scaling dimension (hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT) and entanglement scaling dimension (ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT). Implementing different boundary conditions beyond those considered in this paper should allow one to extract new scaling dimensions, realizing different “permutations” of replicas in the statistical mechanics models describing the measurement induced phase transitions discussed in this paper. Classifying these boundary conditions and understanding their physical meaning in terms of quantum information theoretic quantities remains an important challenge for future work.

Acknowledgements.
We thank M. Gullans, D. Huse, A.C. Potter, J. Wilson and A. Zabalo for insightful discussions and for collaboration on previous works. This work was partially supported by the Abrahams Postdoctoral Fellowship at the Center for Materials Theory Rutgers (A.C.), the Air Force Office of Scientific Research under Grant No. FA9550-21-1-0123 (A.K., R.V.), the Army Research Office Grant No. 79849-PE-H (K.A., J.H.P.) and a Sloan Research Fellowship (J.H.P., R.V.). This work was performed in part at the Aspen Center for Physics, which is supported by National Science Foundation grant PHY-2210452 (J.H.P.). The authors acknowledge the following research computing resources that have contributed to the results reported here: the Open Science Grid Pordes et al. (2007); Sfiligoi et al. (2009), which is supported by the National Science Foundation award 1148698, and the U.S. Department of Energy’s Office of Science.
Refer to caption
Figure 8: Effective central charge in cylindrical geometry: The plot of f𝑓fitalic_f shows linear trend with 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in accordance with the scaling form stated in Eq. 4. We compute ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in cylindrical geometry by the double fitting procedure and we find ceff=0.96(1),0.349(3),subscript𝑐eff0.9610.3493c_{\rm{eff}}=0.96(1),0.349(3),italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.96 ( 1 ) , 0.349 ( 3 ) , and 0.375(5)0.37550.375(5)0.375 ( 5 ) and the extensive bulk parameter f(L=)=0.55718(2),0.131574(6),𝑓𝐿0.5571820.1315746f(L=\infty)=0.55718(2),0.131574(6),italic_f ( italic_L = ∞ ) = 0.55718 ( 2 ) , 0.131574 ( 6 ) , and 0.170469(6)0.17046960.170469(6)0.170469 ( 6 ), obtained to high precision for (a)MOIM, (b)dual Clifford, and (c)Clifford random circuits, respectively. Each data point is averaged over 5×1065superscript1065\times 10^{6}5 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT trajectories.

Appendix A Effective central charge from cylindrical geometry

The free energy density for cylindrical geometry obeys the relation given by Eq. 4. We plot f𝑓fitalic_f vs 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for MOIM, dual Clifford, and Clifford random circuits as shown in Fig 8. The plot shows a linear trend with 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We extract the value of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and f()𝑓f{(\infty)}italic_f ( ∞ ) using the double fitting procedure and we find ceff=0.96(1),0.349(3),0.375(5)subscript𝑐eff0.9610.34930.3755c_{\rm{eff}}=0.96(1),0.349(3),0.375(5)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.96 ( 1 ) , 0.349 ( 3 ) , 0.375 ( 5 ), and the bulk term f(L=)=0.55718(2),0.131574(6),𝑓𝐿0.5571820.1315746f(L=\infty)=0.55718(2),0.131574(6),italic_f ( italic_L = ∞ ) = 0.55718 ( 2 ) , 0.131574 ( 6 ) , and 0.170469(6)0.17046960.170469(6)0.170469 ( 6 ) for MOIM, dual Clifford, and Clifford circuit, respectively.

Appendix B Boundary dissipation in the measurement-only Ising model

Refer to caption
Figure 9: Boundary dissipation in the measurement-only Ising model: (a) The qubits are initialized in a stabilizer state (X1,X2,X3,X4,X5,X6),whereXσxsubscript𝑋1subscript𝑋2subscript𝑋3subscript𝑋4subscript𝑋5subscript𝑋6whereXsuperscript𝜎x(X_{1},X_{2},X_{3},X_{4},X_{5},X_{6})\rm{,~{}where~{}}X\equiv\sigma^{x}( italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT ) , roman_where roman_X ≡ italic_σ start_POSTSUPERSCRIPT roman_x end_POSTSUPERSCRIPT, that undergoes measurement plus dissipation dynamics using the two site measurement operation σizσi+1zsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑧𝑖1\sigma^{z}_{i}\sigma^{z}_{i+1}italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT (blue edge), the single site measurements σixsubscriptsuperscript𝜎𝑥𝑖\sigma^{x}_{i}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (green circles), and the dissipators denoted by purple diamond that act on the right boundary qubit after each time step up to time t=8𝑡8t=8italic_t = 8. In this case we consider maximally mixed depolarizing channel to model dissipation. At the last time step, one artificially measures all sites using σxsuperscript𝜎𝑥\sigma^{x}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT measurement operators. The measurement operations give rise to outcomes that are either random or deterministic. We denote the random outcomes using stars where the red colored stars are the extra random measurements that are absent in the exact same dynamics but with no boundary dissipators. The blue stars are random measurements that are not affected by the boundary dissipation. This particular configuration consists of 6 red and 30 blue stars. (b) Percolation mapping of the configuration shown in (a). The internal and external hulls of boundary clusters (that touch the right boundary) are shown in dark and light blue color, respectively. All other hull of clusters are shown in red. Note that there are 6 boundary hulls, in agreement with eq. (20).
Refer to caption
Figure 10: Boundary spectrum of Clifford circuit: We implement the boundary condition in Fig 1 to obtain the respective boundary spectrum of monitored Clifford circuits. (a) The plot of f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1/L1𝐿1/L1 / italic_L clearly follows the scaling form Eq. 6 (for Lmin{10,12,14,16,18}subscript𝐿min1012141618L_{\rm{min}}\in\{10,12,14,16,18\}italic_L start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ∈ { 10 , 12 , 14 , 16 , 18 }) as shown by the dashed blue line. In the inset, we perform the double fitting procedure to extract ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and f(f|f)ssuperscriptsubscript𝑓conditional𝑓𝑓𝑠f_{(f|f)}^{s}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT. This gives us ceff=0.32(4)subscript𝑐eff0.324c_{\rm{eff}}=0.32(4)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.32 ( 4 ) and f(f|f)s=0.0749(3)superscriptsubscript𝑓conditional𝑓𝑓𝑠0.07493f_{(f|f)}^{s}=-0.0749(3)italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT = - 0.0749 ( 3 ) as shown by the red dashed line. (b) The plot of (f(f|a)f(f|f))Lsubscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓𝐿(f_{(f|a)}-f_{(f|f)})L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ) italic_L scales linearly with 1/L1𝐿1/L1 / italic_L as shown by the dashed blue line, and from fitting the value of slope and intercept we obtain hf|a=0.034(2)subscriptconditional𝑓𝑎0.0342h_{f|a}=0.034(2)italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = 0.034 ( 2 ) and fδs=0.0547(4)subscript𝑓𝛿𝑠0.05474f_{\delta s}=0.0547(4)italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.0547 ( 4 ), shown by the red dashed line in the inset. (c) The plot of SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT vs 1/L1𝐿1/L1 / italic_L clearly follows the trend stated in Eq. 9 and we find ha|b=0.534(2)subscriptconditional𝑎𝑏0.5342h_{a|b}=0.534(2)italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT = 0.534 ( 2 ) and A1=1.9(1)subscript𝐴11.91A_{1}=-1.9(1)italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 1.9 ( 1 ). All the data shown above for the Clifford circuit have pc=0.1596subscript𝑝𝑐0.1596p_{c}=0.1596italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.1596, α=0.61𝛼0.61\alpha=0.61italic_α = 0.61, and L{10,12,14,16,18,20,24,28}𝐿1012141618202428L\in\{10,12,14,16,18,20,24,28\}italic_L ∈ { 10 , 12 , 14 , 16 , 18 , 20 , 24 , 28 }. We sample =5×106absent5superscript106=5\times 10^{6}= 5 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT trajectories to find the average value of free energy density.

Adding boundary dissipation (depolarizers) to the measurement-only Ising model (MOIM) induces additional non-deterministic, random measurements. In turn, the number of such additional random measurements determines the scaling dimension of the BCC scaling dimension hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT using eq. 14 and

F(f|a)F(f|f)=ΔNrandln2,subscript𝐹conditional𝑓𝑎subscript𝐹conditional𝑓𝑓Δsubscript𝑁rand2F_{(f|a)}-F_{(f|f)}=\Delta N_{\rm rand}\ln{2},italic_F start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_F start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT = roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT roman_ln 2 , (19)

where ΔNrand=N¯randDN¯randΔsubscript𝑁randsuperscriptsubscript¯𝑁rand𝐷subscript¯𝑁rand\Delta N_{\rm rand}=\overline{N}_{\rm rand}^{D}-\overline{N}_{\rm rand}roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT = over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT - over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT, with N¯randDsuperscriptsubscript¯𝑁rand𝐷\overline{N}_{\rm rand}^{D}over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT and N¯randsubscript¯𝑁rand\overline{N}_{\rm rand}over¯ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT are the averaged number of random measurements with and without the depolarizers present at the right boundary, respectively. We note that for each realization, ΔNrandΔsubscript𝑁rand\Delta N_{\rm rand}roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT has a simple geometrical meaning in terms of percolation clusters. We find by inspection that

ΔNrand=Nbh,Δsubscript𝑁randsubscript𝑁bh\Delta N_{\rm rand}=N_{\rm bh},roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT roman_bh end_POSTSUBSCRIPT , (20)

where Nbhsubscript𝑁bhN_{\rm bh}italic_N start_POSTSUBSCRIPT roman_bh end_POSTSUBSCRIPT represents the number of hulls of clusters touching the right boundary (where the depolarizers act). We illustrate this relation in Fig. 9. This particular example contains ΔNrand=6Δsubscript𝑁rand6\Delta N_{\rm rand}=6roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT = 6 extra number of random measurements as compared to the case with no dissipator at the right boundary. Those extra measurements are denoted with red stars in Fig. 9(a). Note that these may happen far from right boundary. The corresponding boundary hulls are shown in Fig. 9(b), there are four “external” boundary hulls shown in light blue, and two internal boundary hulls shown in dark blue. The remaining hulls (red) do not contribute to ΔNrandΔsubscript𝑁rand\Delta N_{\rm rand}roman_Δ italic_N start_POSTSUBSCRIPT roman_rand end_POSTSUBSCRIPT. We have checked the validity of eq. (20) numerically for very large circuits.

We now turn to recent boundary CFT results for boundary loop models, that allow us to “tag” boundary hulls and count them at criticality Jacobsen and Saleur (2008). We find numerically that “internal” boundary hulls do not contribute to the universal exponent hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT (not shown), while the “external” boundary hulls can be counted using Ref. Jacobsen and Saleur, 2008:

Nbhexternaltπ2L34π,similar-tosubscriptsuperscript𝑁externalbh𝑡𝜋2𝐿34𝜋\frac{N^{\rm external}_{\rm bh}}{t}\sim\frac{\pi}{2L}\frac{\sqrt{3}}{4\pi},divide start_ARG italic_N start_POSTSUPERSCRIPT roman_external end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_bh end_POSTSUBSCRIPT end_ARG start_ARG italic_t end_ARG ∼ divide start_ARG italic_π end_ARG start_ARG 2 italic_L end_ARG divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 4 italic_π end_ARG , (21)

ignoring non-universal, 𝒪(1)𝒪1{\cal O}(1)caligraphic_O ( 1 ) contributions, and where the factor of 2L2𝐿2L2 italic_L is the number of sites in the loop (or Majorana) language. Using eqs. (14) and (19), we identify the scaling dimension

hf|a=38πln2.subscriptconditional𝑓𝑎38𝜋2h_{f|a}=\frac{\sqrt{3}}{8\pi}\ln 2.italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT = divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 8 italic_π end_ARG roman_ln 2 . (22)

Appendix C Clifford random circuits

The dynamics of Clifford random circuit can be expressed using the transfer matrix approach discussed in section II. We thus use the scaling form of free energy and SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT to compute the boundary conformal properties for the Clifford MIPT.

ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT.—

We will first start by discussing the calculation of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in cylindrical geometry. In fig 8(c) the plot of f𝑓fitalic_f shows linear trend with 1/L21superscript𝐿21/L^{2}1 / italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in accordance with the CFT result stated in Eq. 4. We note that the double fit method gives ceff=0.375(5)subscript𝑐eff0.3755c_{\rm{eff}}=0.375(5)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.375 ( 5 ) which is in agreement with the previously known result Zabalo et al. (2022) and the extensive bulk parameter f(L=)=0.170469(6)𝑓𝐿0.1704696f(L=\infty)=0.170469(6)italic_f ( italic_L = ∞ ) = 0.170469 ( 6 ). We now implement Clifford circuit with free boundary ends where we find ceff=0.32(4)subscript𝑐eff0.324c_{\rm{eff}}=0.32(4)italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.32 ( 4 ) from the plot between f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1/L1𝐿1/L1 / italic_L as shown in fig 5(a). The plot respects the scaling form given in Eq. 6 and the value of ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT validate the result for dual Clifford circuit as noted in Table 1. The extracted values from the plot between f(f|f)subscript𝑓conditional𝑓𝑓f_{(f|f)}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT and 1L1𝐿\frac{1}{L}divide start_ARG 1 end_ARG start_ARG italic_L end_ARG gives leading order correction to Eq. 6 since we have ceff(L)=0.326.608L2subscript𝑐eff𝐿0.326.608superscript𝐿2c_{\rm{eff}}(L)=0.32-\frac{6.608}{L^{2}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_L ) = 0.32 - divide start_ARG 6.608 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and f(f|f)s(L)=0.0750.042L2superscriptsubscript𝑓conditional𝑓𝑓𝑠𝐿0.0750.042superscript𝐿2f_{(f|f)}^{s}(L)=-0.075-\frac{0.042}{L^{2}}italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( italic_L ) = - 0.075 - divide start_ARG 0.042 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG using the double fitting procedure.

hf|asubscriptconditional𝑓𝑎h_{f|a}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT.—

The plot of (f(f|a)f(f|f))Lsubscript𝑓conditional𝑓𝑎subscript𝑓conditional𝑓𝑓𝐿(f_{(f|a)}-f_{(f|f)})L( italic_f start_POSTSUBSCRIPT ( italic_f | italic_a ) end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT ( italic_f | italic_f ) end_POSTSUBSCRIPT ) italic_L vs 1L21superscript𝐿2\frac{1}{L^{2}}divide start_ARG 1 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG displays a linear trend as shown in 10(b). Now using the double fitting procedure we find the boundary scaling dimension hf|a(L)=0.0340.442L2subscriptconditional𝑓𝑎𝐿0.0340.442superscript𝐿2h_{f|a}(L)=0.034-\frac{0.442}{L^{2}}italic_h start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT ( italic_L ) = 0.034 - divide start_ARG 0.442 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG and the non-universal surface term fδs=0.055+0.068L2subscript𝑓𝛿𝑠0.0550.068superscript𝐿2f_{\delta s}=0.055+\frac{0.068}{L^{2}}italic_f start_POSTSUBSCRIPT italic_δ italic_s end_POSTSUBSCRIPT = 0.055 + divide start_ARG 0.068 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.

ha|bsubscriptconditional𝑎𝑏h_{a|b}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT.—

The plot in fig 10(c) clearly respects the scaling form of SMIEsubscript𝑆MIES_{\rm{MIE}}italic_S start_POSTSUBSCRIPT roman_MIE end_POSTSUBSCRIPT as given in Eq. 9 and we extract the entanglement scaling dimension ha|b(L)=0.5340.517L2subscriptconditional𝑎𝑏𝐿0.5340.517superscript𝐿2h_{a|b}(L)=0.534-\frac{0.517}{L^{2}}italic_h start_POSTSUBSCRIPT italic_a | italic_b end_POSTSUBSCRIPT ( italic_L ) = 0.534 - divide start_ARG 0.517 end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG which agrees with known results for Clifford random circuits Li et al. (2021b).

C.1 Saturation time

We investigate the dynamics of free energy density to measure the saturation time with and without depolarizers for Clifford random circuit as shown in Fig 11. We initialize the circuit with a random product or a random stabilizer state and then compute free energy density by averaging over 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT sampled trajectories. One clearly notice saturation above the equilibration time 4Lsimilar-toabsent4𝐿\sim 4L∼ 4 italic_L and the average in fig 11 represents the mean of free energy associated with the two types of initial state.

Refer to caption
Figure 11: Saturation time: Evolution of free energy density for Clifford circuit with (a) free boundary ends and (b)with depolarizing channel at the right boundary end. It realizes saturation near the equilibration time 4Lsimilar-toabsent4𝐿\sim 4L∼ 4 italic_L both for random product (blue) and random stabilizer initial states(orange). The average of two states together is shown in green. Following are the parameters for both the plots: L=16𝐿16L=16italic_L = 16, pc=0.1596subscript𝑝𝑐0.1596p_{c}=0.1596italic_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.1596, and number of samples =105absentsuperscript105=10^{5}= 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT.

Appendix D Methods of Monte Carlo Sampling

A quantum system undergoing decoherence can be described by the map ρ(t+dt)=kM^kρM^k𝜌𝑡𝑑𝑡subscript𝑘subscript^𝑀𝑘𝜌subscriptsuperscript^𝑀𝑘\rho(t+dt)=\sum_{k}\hat{M}_{k}\rho\hat{M}^{\dagger}_{k}italic_ρ ( italic_t + italic_d italic_t ) = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_ρ over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, where the set of Krauss Operators {M^k}subscript^𝑀𝑘\{\hat{M}_{k}\}{ over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } satisfy kM^kM^k=Isubscript𝑘subscriptsuperscript^𝑀𝑘subscript^𝑀𝑘𝐼\sum_{k}\hat{M}^{\dagger}_{k}\hat{M}_{k}=I∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_I, over the set of possible measurement outcomes k𝑘kitalic_k. The choice of Mksubscript𝑀𝑘M_{k}italic_M start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is not necessarily unique. If the state were initially pure, then after a measurement described by M^ksubscript^𝑀𝑘\hat{M}_{k}over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, |ψM^k|ψpkket𝜓subscript^𝑀𝑘ket𝜓subscript𝑝𝑘|\psi\rangle\rightarrow\frac{\hat{M}_{k}|\psi\rangle}{\sqrt{p_{k}}}| italic_ψ ⟩ → divide start_ARG over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_ψ ⟩ end_ARG start_ARG square-root start_ARG italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_ARG, with probability pk=Tr(M^k|ψψ|M^kp_{k}=\mathrm{Tr}({\hat{M}_{k}|\psi\rangle\langle\psi|}\hat{M}^{\dagger}_{k}italic_p start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = roman_Tr ( over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_ψ ⟩ ⟨ italic_ψ | over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT). The Monte Carlo method is implemented by starting from a pure initial state, and evolving |ψket𝜓|\psi\rangle| italic_ψ ⟩ with M^ksubscript^𝑀𝑘\hat{M}_{k}over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT with probability ψ|MkMk|ψ.quantum-operator-product𝜓subscriptsuperscript𝑀𝑘subscript𝑀𝑘𝜓\langle\psi|M^{\dagger}_{k}M_{k}|\psi\rangle.⟨ italic_ψ | italic_M start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_ψ ⟩ . The dephasing channel can be written with k=2𝑘2k=2italic_k = 2 Krauss operators: M^0=Psubscript^𝑀0subscript𝑃\hat{M}_{0}=P_{\uparrow}over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT and M^1=Psubscript^𝑀1subscript𝑃\hat{M}_{1}=P_{\downarrow}over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , where

ρ(t+dt)=M^0ρM^0+M^1ρM^1,𝜌𝑡𝑑𝑡subscript^𝑀0𝜌superscriptsubscript^𝑀0subscript^𝑀1𝜌superscriptsubscript^𝑀1\rho(t+dt)=\hat{M}_{0}\rho\hat{M}_{0}^{{\dagger}}+\hat{M}_{1}\rho\hat{M}_{1}^{% {\dagger}},italic_ρ ( italic_t + italic_d italic_t ) = over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , (23)

or alternatively with Krauss operators

M^0(dt)=𝕀c1c12dt=1pdI,subscriptsuperscript^𝑀0𝑑𝑡𝕀superscriptsubscript𝑐1subscript𝑐12𝑑𝑡1subscript𝑝𝑑𝐼\displaystyle\hat{M}^{{}^{\prime}}_{0}(dt)=\mathbb{I}-\frac{c_{1}^{{\dagger}}c% _{1}}{2}dt=\sqrt{1-p_{d}}I,over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_d italic_t ) = blackboard_I - divide start_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_d italic_t = square-root start_ARG 1 - italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG italic_I ,
M^1(dt)=dtc1=pdσzsubscriptsuperscript^𝑀1𝑑𝑡𝑑𝑡superscriptsubscript𝑐1subscript𝑝𝑑subscript𝜎𝑧\displaystyle\hat{M}^{{}^{\prime}}_{1}(dt)=\sqrt{dt}c_{1}^{{\dagger}}=\sqrt{p_% {d}}\sigma_{z}over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_d italic_t ) = square-root start_ARG italic_d italic_t end_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT = square-root start_ARG italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (24)

where c^=1pddtσz^𝑐1subscript𝑝𝑑𝑑𝑡subscript𝜎𝑧\hat{c}=\sqrt{\frac{1-p_{d}}{dt}}\sigma_{z}over^ start_ARG italic_c end_ARG = square-root start_ARG divide start_ARG 1 - italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the Lindblad operator for a dephasing channel along the z-axis and pd=1/2subscript𝑝𝑑12p_{d}=1/2italic_p start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 1 / 2. In a given Monte Carlo sample of the dissipator, |ψket𝜓|\psi\rangle| italic_ψ ⟩ undergoes a “jump”, or evolution with c^^𝑐\hat{c}over^ start_ARG italic_c end_ARG, with probability pjump=Tr[M^1ρM^1]subscript𝑝jumpTrdelimited-[]subscriptsuperscript^𝑀1𝜌subscriptsuperscript^superscript𝑀1p_{\mathrm{jump}}=\mathrm{Tr}[\hat{M}^{{}^{\prime}}_{1}\rho\hat{M^{{}^{\prime}% }}^{\dagger}_{1}]italic_p start_POSTSUBSCRIPT roman_jump end_POSTSUBSCRIPT = roman_Tr [ over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ over^ start_ARG italic_M start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ]. In the simulation, a random number r𝑟ritalic_r, chosen from the unit interval, and if r<pjump𝑟subscript𝑝jumpr<p_{\mathrm{jump}}italic_r < italic_p start_POSTSUBSCRIPT roman_jump end_POSTSUBSCRIPT, the state vector evolves as M^1(dt)|ψ(t)ψ(t)|M^1(dt)M^1(dt)|ψ(t).subscriptsuperscript^𝑀1𝑑𝑡ket𝜓𝑡quantum-operator-product𝜓𝑡subscriptsuperscript^𝑀1𝑑𝑡subscriptsuperscript^𝑀1𝑑𝑡𝜓𝑡\frac{\hat{M}^{\prime}_{1}(dt)|\psi(t)\rangle}{\sqrt{\langle\psi(t)|\hat{M}^{% \prime{\dagger}}_{1}(dt)\hat{M}^{\prime}_{1}(dt)|\psi(t)\rangle}}.divide start_ARG over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_d italic_t ) | italic_ψ ( italic_t ) ⟩ end_ARG start_ARG square-root start_ARG ⟨ italic_ψ ( italic_t ) | over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT ′ † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_d italic_t ) over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_d italic_t ) | italic_ψ ( italic_t ) ⟩ end_ARG end_ARG . Whereas, if no jump occurs, |ψ(t)ket𝜓𝑡|\psi(t)\rangle| italic_ψ ( italic_t ) ⟩ is unchanged.

Refer to caption
Figure 12: (a) Ff|asubscript𝐹conditional𝑓𝑎F_{f|a}italic_F start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT extrapolated to various values of τ𝜏\tauitalic_τ. (b) ff|asubscript𝑓conditional𝑓𝑎f_{f|a}italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT as a function of 1/τ1𝜏1/\tau1 / italic_τ, where τ𝜏\tauitalic_τ is the number of Monte Carlo iterations used to calculate ff|asubscript𝑓conditional𝑓𝑎f_{f|a}italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT. The dashed blue line is extrapolated fit, which is also shown in the inset. The dashed red line in the main plot is the full density matrix data. Inset: ff|asubscript𝑓conditional𝑓𝑎f_{f|a}italic_f start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT computed from Ff|asubscript𝐹conditional𝑓𝑎F_{f|a}italic_F start_POSTSUBSCRIPT italic_f | italic_a end_POSTSUBSCRIPT extrapolated to τ𝜏\tauitalic_τ. The black line is computed from the free energy density without extrapolation, starting from τitsubscript𝜏𝑖𝑡\tau_{it}italic_τ start_POSTSUBSCRIPT italic_i italic_t end_POSTSUBSCRIPT = 300 to τitsubscript𝜏𝑖𝑡\tau_{it}italic_τ start_POSTSUBSCRIPT italic_i italic_t end_POSTSUBSCRIPT = 600. The blue line is the extrapolation to τit=subscript𝜏𝑖𝑡\tau_{it}=\inftyitalic_τ start_POSTSUBSCRIPT italic_i italic_t end_POSTSUBSCRIPT = ∞ for L=6𝐿6L=6italic_L = 6 and L=8𝐿8L=8italic_L = 8. For L=6𝐿6L=6italic_L = 6, the full density matrix yields f=0.07519188(1)𝑓0.075191881f=\mathrm{0.07519188(1)}italic_f = 0.07519188 ( 1 ), and f(τ=)=0.07514(2)𝑓𝜏0.075142f(\tau=\infty)=0.07514(2)italic_f ( italic_τ = ∞ ) = 0.07514 ( 2 ) from the extrapolation method. For L=8𝐿8L=8italic_L = 8, f=0.07865890(1)𝑓0.078658901f=\mathrm{0.07865890(1)}italic_f = 0.07865890 ( 1 ), and f(τ=)=0.07857(2)𝑓𝜏0.078572f(\tau=\infty)=\mathrm{0.07857(2)}italic_f ( italic_τ = ∞ ) = 0.07857 ( 2 ).

Each circuit realization is specified by a set of two-site random unitary gates U={Ut,ij}𝑈subscript𝑈𝑡𝑖𝑗U=\{U_{t,ij}\}italic_U = { italic_U start_POSTSUBSCRIPT italic_t , italic_i italic_j end_POSTSUBSCRIPT }, the space-time measurement locations X𝑋\vec{X}over→ start_ARG italic_X end_ARG, and the measurement trajectory 𝐦𝐦{\bf m}bold_m yielding the outcomes of the measurements. After T𝑇Titalic_T timesteps, the circuit yields the unnormalized state:

ρ𝐦=ρ𝐦(U,X,T)=𝒟{PTUT𝒟{P2U2P1U1ρU1P1U2P2}UTPT}.subscript𝜌𝐦subscript𝜌𝐦𝑈𝑋𝑇𝒟subscript𝑃𝑇subscript𝑈𝑇𝒟subscript𝑃2subscript𝑈2subscript𝑃1subscript𝑈1𝜌superscriptsubscript𝑈1subscript𝑃1superscriptsubscript𝑈2subscript𝑃2subscriptsuperscript𝑈𝑇subscript𝑃𝑇\rho_{{\bf m}}=\rho_{{\bf m}}(U,\vec{X},T)=\mathcal{D}\{P_{T}U_{T}...\mathcal{% D}\{P_{2}U_{2}P_{1}U_{1}\rho U_{1}^{{\dagger}}P_{1}U_{2}^{{\dagger}}P_{2}\}...% U^{{\dagger}}_{T}P_{T}\}.italic_ρ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = italic_ρ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_U , over→ start_ARG italic_X end_ARG , italic_T ) = caligraphic_D { italic_P start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT … caligraphic_D { italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT } … italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT } . (25)

One timestep consists of a layer of either even or odd gates and measurements. The probability of a meausurement trajectory 𝐦𝐦{\bf m}bold_m is p𝐦(U,X,T)=Tr(ρ𝐦(U,X,T))subscript𝑝𝐦𝑈𝑋𝑇Trsubscript𝜌𝐦𝑈𝑋𝑇p_{{\bf m}}(U,\vec{X},T)=\mathrm{Tr}(\rho_{{\bf m}}(U,\vec{X},T))italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_U , over→ start_ARG italic_X end_ARG , italic_T ) = roman_Tr ( italic_ρ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_U , over→ start_ARG italic_X end_ARG , italic_T ) ). The dephasing channel can be rewritten as

ρ𝐦=iM^iT{PTUTM^i2{P2U2P1U1ρU1P1U2P2}M^i2UTPT}M^iT,subscript𝜌𝐦subscript𝑖subscript^𝑀subscript𝑖𝑇subscript𝑃𝑇subscript𝑈𝑇subscript^𝑀subscript𝑖2subscript𝑃2subscript𝑈2subscript𝑃1subscript𝑈1𝜌superscriptsubscript𝑈1subscript𝑃1superscriptsubscript𝑈2subscript𝑃2subscript^𝑀subscript𝑖2subscriptsuperscript𝑈𝑇subscript𝑃𝑇subscript^𝑀subscript𝑖𝑇\rho_{{\bf m}}=\sum_{\vec{i}}\hat{M}_{i_{T}}\{P_{T}U_{T}...\hat{M}_{i_{2}}\{P_% {2}U_{2}P_{1}U_{1}\rho U_{1}^{{\dagger}}P_{1}U_{2}^{{\dagger}}P_{2}\}\hat{M}_{% i_{2}}...U^{{\dagger}}_{T}P_{T}\}\hat{M}_{i_{T}},italic_ρ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT over→ start_ARG italic_i end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT { italic_P start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT … over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT { italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT } over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT … italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT } over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT , (26)

where M^ijsubscript^𝑀subscript𝑖𝑗\hat{M}_{i_{j}}over^ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT is an operator chosen from either Krauss representation and we sum over all 2T/2superscript2𝑇22^{T/2}2 start_POSTSUPERSCRIPT italic_T / 2 end_POSTSUPERSCRIPT possible dephasing trajectories labelled by i={i2,,iT}𝑖subscript𝑖2subscript𝑖𝑇\vec{i}=\{i_{2},...,i_{T}\}over→ start_ARG italic_i end_ARG = { italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT } for fixed (U,X,𝐦)𝑈𝑋𝐦(U,\vec{X},{\bf m})( italic_U , over→ start_ARG italic_X end_ARG , bold_m ). We can now write the probability of a given Lindblad trajectory described by the specific Krauss operators i2,i4,,iT={Mi2,Mi4,,MiT}subscriptsubscript𝑖2subscript𝑖4subscript𝑖𝑇subscriptsuperscript𝑀subscript𝑖2subscriptsuperscript𝑀subscript𝑖4subscriptsuperscript𝑀subscript𝑖𝑇\mathcal{M}_{i_{2},i_{4},...,i_{T}}=\{M^{{}^{\prime}}_{i_{2}},M^{{}^{\prime}}_% {i_{4}},...,M^{{}^{\prime}}_{i_{T}}\}caligraphic_M start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT = { italic_M start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_M start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_M start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT }, where it{0,1}subscript𝑖𝑡01i_{t}\in\{0,1\}italic_i start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ∈ { 0 , 1 } labels the Krauss operator for the dephasing channel at the tthsubscript𝑡𝑡t_{th}italic_t start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT layer. For each (U,X,𝐦)𝑈𝑋𝐦(U,\vec{X},{\bf m})( italic_U , over→ start_ARG italic_X end_ARG , bold_m ), we apply M^0subscriptsuperscript^𝑀0\hat{M}^{{}^{\prime}}_{0}over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT or M^1subscriptsuperscript^𝑀1\hat{M}^{{}^{\prime}}_{1}over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT with equal probability on x=L𝑥𝐿x=Litalic_x = italic_L at each timestep. After T𝑇Titalic_T timesteps we have,

pi2,i4,,iT=Tr{M^iTP2TU2T{M^i2P2U2P1U1ρU1P1U2P2M^i2}U2TP2TM^iT}.subscript𝑝subscript𝑖2subscript𝑖4subscript𝑖𝑇Trsubscriptsuperscript^𝑀subscript𝑖𝑇subscript𝑃2𝑇subscript𝑈2𝑇subscriptsuperscript^𝑀subscript𝑖2subscript𝑃2subscript𝑈2subscript𝑃1subscript𝑈1𝜌subscriptsuperscript𝑈1subscript𝑃1subscriptsuperscript𝑈2subscript𝑃2subscriptsuperscript^𝑀subscript𝑖2subscriptsuperscript𝑈2𝑇subscript𝑃2𝑇subscriptsuperscript^𝑀subscript𝑖𝑇p_{{i_{2}},i_{4},...,i_{T}}=\mathrm{Tr}\{\hat{M}^{{}^{\prime}}_{i_{T}}P_{2T}U_% {2T}...\{\hat{M}^{{}^{\prime}}_{{i}_{2}}P_{2}U_{2}P_{1}U_{1}\rho U^{\dagger}_{% 1}P_{1}U^{\dagger}_{2}P_{2}\hat{M}^{{}^{\prime}}_{{i}_{2}}\}...U^{\dagger}_{2T% }P_{2T}\hat{M}^{{}^{\prime}}_{i_{T}}\}.italic_p start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT = roman_Tr { over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 2 italic_T end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 italic_T end_POSTSUBSCRIPT … { over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT } … italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 italic_T end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 2 italic_T end_POSTSUBSCRIPT over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT } . (27)

We compute pi2,i4,,iTsubscript𝑝subscript𝑖2subscript𝑖4subscript𝑖𝑇p_{i_{2},i_{4},...,i_{T}}italic_p start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT for several different trajectories and estimate the Born probability after averaging over all possible dissipator outcomes p𝐦subscript𝑝𝐦p_{{\bf m}}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT

p𝐦=[pi2,i4,,iT]MCsubscript𝑝𝐦subscriptdelimited-[]subscript𝑝subscript𝑖2subscript𝑖4subscript𝑖𝑇𝑀𝐶p_{{\bf m}}=\left[p_{i_{2},i_{4},...,i_{T}}\right]_{MC}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = [ italic_p start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT (28)

where []MCsubscriptdelimited-[]𝑀𝐶[\dots]_{MC}[ … ] start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT denotes a Monte Carlo average over dissipator outcomes for many samples.

D.1 Extrapolation Method

In this subsection, we describe our numerical method of Monte Carlo sampling from the full density density matrix to approximate the free energy. For a fixed (U,X)𝑈𝑋(U,\vec{X})( italic_U , over→ start_ARG italic_X end_ARG ), the free energy of the measurement record, F𝐹Fitalic_F, is the average of the logarithm of the probability of a given trajectory:

F=𝐦p𝐦lnp𝐦=𝐦lnp𝐦.𝐹subscript𝐦subscript𝑝𝐦subscript𝑝𝐦subscript𝐦delimited-⟨⟩subscript𝑝𝐦F=-\sum_{{\bf m}}p_{{\bf m}}\ln p_{{\bf m}}=-\sum_{{\bf m}}\langle\ln p_{{\bf m% }}\rangle.italic_F = - ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT roman_ln italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ⟨ roman_ln italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ⟩ . (29)

We perform the Monte Carlo sampling over the dissipator outcomes for a fixed projective measurement trajectory 𝐦𝐦{\bf m}bold_m and fixed (U,X)𝑈𝑋(U,\vec{X})( italic_U , over→ start_ARG italic_X end_ARG ). The probability of a given Monte Carlo trajectory i𝑖iitalic_i, consisting of a fixed sequence of Monte Carlo propagators from the dissipator, is p𝐦(i).subscript𝑝𝐦𝑖p_{{\bf m}}{(i)}.italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_i ) . Obtaining many samples i𝑖iitalic_i allows us to approximate p𝐦=[p𝐦(i)]MCsubscript𝑝𝐦subscriptdelimited-[]subscript𝑝𝐦𝑖𝑀𝐶p_{{\bf m}}=[p_{{\bf m}}(i)]_{MC}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT = [ italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_i ) ] start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT for a fixed set of gates and measurements.

The accuracy of the sampling method depends on the number of Monte Carlo samples taken, due to the statistics of the dissipator outcomes. Due to the large number of Monte Carlo trajectories with high p𝐦subscript𝑝𝐦p_{{\bf m}}italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT, p𝐦(i)delimited-⟨⟩subscript𝑝𝐦𝑖\langle p_{{\bf m}}(i)\rangle⟨ italic_p start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT ( italic_i ) ⟩ decreases as the number of Monte Carlo trajectories increases.

To estimate the entropy of the measurement record, we record the free energy as a function of the number of Monte Carlo samples (denoted as τ𝜏\tauitalic_τ) at each time: F(t;τ).𝐹𝑡𝜏F(t;\tau).italic_F ( italic_t ; italic_τ ) . To determine the free energy when τ𝜏\tau\rightarrow\inftyitalic_τ → ∞, we compute the free energy averaged over multiple circuit realizations and outcomes as a function of the number of Monte Carlo iterations. We extrapolate to τ𝜏\tau\rightarrow\inftyitalic_τ → ∞ from τmin=300superscript𝜏min300\tau^{\mathrm{min}}=300italic_τ start_POSTSUPERSCRIPT roman_min end_POSTSUPERSCRIPT = 300 to τmax=600superscript𝜏max600\tau^{\mathrm{max}}=600italic_τ start_POSTSUPERSCRIPT roman_max end_POSTSUPERSCRIPT = 600, where F𝐹Fitalic_F exhibits a leading 1/τ1𝜏1/\tau1 / italic_τ dependence for all system sizes studied, as depicted in Fig. 12. We obtain an estimate for F(t;τ)𝐹𝑡𝜏F(t;\tau\rightarrow\infty)italic_F ( italic_t ; italic_τ → ∞ ) from the intercept of the extrapolation as τ𝜏\tau\rightarrow\inftyitalic_τ → ∞ using a fourth order polynomial in 1/τ1𝜏1/\tau1 / italic_τ. From the intercept of the extrapolation we estimate F(t;τ)𝐹𝑡𝜏F(t;\tau\rightarrow\infty)italic_F ( italic_t ; italic_τ → ∞ ). The accuracy of the approach is exemplified in Fig. 12 and discussed in more detail in its caption.

To compute the error bars, we record f𝑓fitalic_f from individual trajectories, and calculate the bootstrap standard error σ𝜎\sigmaitalic_σ for 500500500500 samples for random Haar and product initial states. The errors are combined using σ=12σproduct2+σHaar2𝜎12superscriptsubscript𝜎product2superscriptsubscript𝜎Haar2\sigma=\frac{1}{2}\sqrt{\sigma_{\rm{product}}^{2}+\sigma_{\rm{Haar}}^{2}}italic_σ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG square-root start_ARG italic_σ start_POSTSUBSCRIPT roman_product end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT roman_Haar end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, for each f(L)𝑓𝐿f(L)italic_f ( italic_L ) point. The size of the error bar in ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and fssubscript𝑓𝑠f_{s}italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is determined by the range of possible ceffsubscript𝑐effc_{\rm{eff}}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and fssubscript𝑓𝑠f_{s}italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for f𝑓fitalic_f varying within σ𝜎\sigmaitalic_σ.

References

  • Li et al. (2018) Y. Li, X. Chen,  and M. P. A. Fisher, Phys. Rev. B 98, 205136 (2018).
  • Li et al. (2019) Y. Li, X. Chen,  and M. P. A. Fisher, Phys. Rev. B 100, 134306 (2019).
  • Skinner et al. (2019) B. Skinner, J. Ruhman,  and A. Nahum, Physical Review X 9, 031009 (2019).
  • Chan et al. (2019) A. Chan, R. M. Nandkishore, M. Pretko,  and G. Smith, Physical Review B 99, 224307 (2019).
  • Choi et al. (2020) S. Choi, Y. Bao, X.-L. Qi,  and E. Altman, Phys. Rev. Lett. 125, 030505 (2020).
  • Potter and Vasseur (2022) A. C. Potter and R. Vasseur, in Entanglement in Spin Chains: From Theory to Quantum Technology Applications (Springer, 2022) pp. 211–249.
  • Fisher et al. (2023) M. P. A. Fisher, V. Khemani, A. Nahum,  and S. Vijay, Annual Review of Condensed Matter Physics 14, 335 (2023).
  • Cao et al. (2019) X. Cao, A. Tilloy,  and A. De Luca, SciPost Physics 7, 024 (2019).
  • Szyniszewski et al. (2019) M. Szyniszewski, A. Romito,  and H. Schomerus, Phys. Rev. B 100, 064204 (2019).
  • Lunt and Pal (2020) O. Lunt and A. Pal, Phys. Rev. Res. 2, 043072 (2020).
  • Goto and Danshita (2020) S. Goto and I. Danshita, Phys. Rev. A 102, 033316 (2020).
  • Tang and Zhu (2020) Q. Tang and W. Zhu, Phys. Rev. Res. 2, 013022 (2020).
  • Iaconis et al. (2020) J. Iaconis, A. Lucas,  and X. Chen, Phys. Rev. B 102, 224311 (2020).
  • Turkeshi et al. (2020) X. Turkeshi, R. Fazio,  and M. Dalmonte, Phys. Rev. B 102, 014315 (2020).
  • Zhang et al. (2020) L. Zhang, J. A. Reyes, S. Kourtis, C. Chamon, E. R. Mucciolo,  and A. E. Ruckenstein, Phys. Rev. B 101, 235104 (2020).
  • Szyniszewski et al. (2020) M. Szyniszewski, A. Romito,  and H. Schomerus, Phys. Rev. Lett. 125, 210602 (2020).
  • Fuji and Ashida (2020) Y. Fuji and Y. Ashida, Phys. Rev. B 102, 054302 (2020).
  • Rossini and Vicari (2020) D. Rossini and E. Vicari, Phys. Rev. B 102, 035119 (2020).
  • Vijay (2020) S. Vijay, arXiv preprint arXiv:2005.03052  (2020).
  • Turkeshi et al. (2021) X. Turkeshi, A. Biella, R. Fazio, M. Dalmonte,  and M. Schiró, Phys. Rev. B 103, 224210 (2021).
  • Chen (2021) X. Chen, arXiv preprint arXiv:2110.12230  (2021).
  • Li and Fisher (2021a) Y. Li and M. P. A. Fisher, arXiv preprint arXiv:2108.04274  (2021a).
  • Lavasani et al. (2021a) A. Lavasani, Y. Alavirad,  and M. Barkeshli, Phys. Rev. Lett. 127, 235701 (2021a).
  • Van Regemortel et al. (2021) M. Van Regemortel, Z.-P. Cian, A. Seif, H. Dehghani,  and M. Hafezi, Phys. Rev. Lett. 126, 123604 (2021).
  • Lunt et al. (2021) O. Lunt, M. Szyniszewski,  and A. Pal, Phys. Rev. B 104, 155111 (2021).
  • Alberton et al. (2021) O. Alberton, M. Buchhold,  and S. Diehl, Phys. Rev. Lett. 126, 170602 (2021).
  • Nahum et al. (2021) A. Nahum, S. Roy, B. Skinner,  and J. Ruhman, PRX Quantum 2, 010352 (2021).
  • Han and Chen (2022) Y. Han and X. Chen, Phys. Rev. B 105, 064306 (2022).
  • Sierant et al. (2022a) P. Sierant, G. Chiriacò, F. M. Surace, S. Sharma, X. Turkeshi, M. Dalmonte, R. Fazio,  and G. Pagano, Quantum 6, 638 (2022a).
  • Sharma et al. (2022) S. Sharma, X. Turkeshi, R. Fazio,  and M. Dalmonte, SciPost Physics Core 5, 023 (2022).
  • Feng et al. (2022) X. Feng, B. Skinner,  and A. Nahum, arXiv preprint arXiv:2210.07264  (2022).
  • Iadecola et al. (2022) T. Iadecola, S. Ganeshan, J. Pixley,  and J. H. Wilson, arXiv preprint arXiv:2207.12415  (2022).
  • Buchhold et al. (2022) M. Buchhold, T. Mueller,  and S. Diehl, arXiv preprint arXiv:2208.10506  (2022).
  • Sierant and Turkeshi (2023a) P. Sierant and X. Turkeshi, Phys. Rev. Lett. 130, 120402 (2023a).
  • O’Dea et al. (2022) N. O’Dea, A. Morningstar, S. Gopalakrishnan,  and V. Khemani, arXiv preprint arXiv:2211.12526  (2022).
  • Piroli et al. (2023) L. Piroli, Y. Li, R. Vasseur,  and A. Nahum, Phys. Rev. B 107, 224303 (2023).
  • Sierant and Turkeshi (2023b) P. Sierant and X. Turkeshi, arXiv preprint arXiv:2308.13384  (2023b).
  • LeMaire et al. (2023) C. LeMaire, A. A. Allocca, J. Pixley, T. Iadecola,  and J. H. Wilson, arXiv preprint arXiv:2309.04520  (2023).
  • Ravindranath et al. (2023a) V. Ravindranath, Y. Han, Z.-C. Yang,  and X. Chen, Phys. Rev. B 108, L041103 (2023a).
  • Jian et al. (2023) C.-M. Jian, H. Shapourian, B. Bauer,  and A. W. W. Ludwig, arXiv preprint arXiv:2302.09094  (2023).
  • Fava et al. (2023) M. Fava, L. Piroli, T. Swann, D. Bernard,  and A. Nahum, arXiv preprint arXiv:2302.12820  (2023).
  • Ravindranath et al. (2023b) V. Ravindranath, Z.-C. Yang,  and X. Chen, arXiv preprint arXiv:2306.16595  (2023b).
  • Roser et al. (2023) F. Roser, H. P. Büchler,  and N. Lang, Phys. Rev. B 107, 214201 (2023).
  • Coppola et al. (2022) M. Coppola, E. Tirrito, D. Karevski,  and M. Collura, Phys. Rev. B 105, 094303 (2022).
  • Tirrito et al. (2023) E. Tirrito, A. Santini, R. Fazio,  and M. Collura, SciPost Physics 15, 096 (2023).
  • Gullans and Huse (2020a) M. J. Gullans and D. A. Huse, Phys. Rev. X 10, 041020 (2020a).
  • Li and Fisher (2021b) Y. Li and M. P. A. Fisher, Phys. Rev. B 103, 104306 (2021b).
  • Lovas et al. (2023) I. Lovas, U. Agrawal,  and S. Vijay, arXiv preprint arXiv:2304.02664  (2023).
  • Agrawal et al. (2022) U. Agrawal, A. Zabalo, K. Chen, J. H. Wilson, A. C. Potter, J. H. Pixley, S. Gopalakrishnan,  and R. Vasseur, Phys. Rev. X 12, 041002 (2022).
  • Barratt et al. (2022a) F. Barratt, U. Agrawal, A. C. Potter, S. Gopalakrishnan,  and R. Vasseur, Phys. Rev. Lett. 129, 200602 (2022a).
  • Barratt et al. (2022b) F. Barratt, U. Agrawal, S. Gopalakrishnan, D. A. Huse, R. Vasseur,  and A. C. Potter, Phys. Rev. Lett. 129, 120604 (2022b).
  • Majidy et al. (2023) S. Majidy, U. Agrawal, S. Gopalakrishnan, A. C. Potter, R. Vasseur,  and N. Y. Halpern, arXiv preprint arXiv:2305.13356  (2023).
  • Li et al. (2023a) Y. Li, Y. Zou, P. Glorioso, E. Altman,  and M. P. A. Fisher, Physical Review Letters 130, 220404 (2023a).
  • Tikhanovskaya et al. (2023) M. Tikhanovskaya, A. Lavasani, M. P. A. Fisher,  and S. Vijay, arXiv preprint arXiv:2306.00058  (2023).
  • Ippoliti and Khemani (2023) M. Ippoliti and V. Khemani, arXiv preprint arXiv:2307.15011  (2023).
  • Bao et al. (2020) Y. Bao, S. Choi,  and E. Altman, Phys. Rev. B 101, 104301 (2020).
  • Jian et al. (2020) C.-M. Jian, Y.-Z. You, R. Vasseur,  and A. W. W. Ludwig, Phys. Rev. B 101, 104302 (2020).
  • Li et al. (2021a) Y. Li, R. Vasseur, M. P. A. Fisher,  and A. W. W. Ludwig, arXiv preprint arXiv:2110.02988  (2021a).
  • Li et al. (2023b) Y. Li, S. Vijay,  and M. P. A. Fisher, PRX Quantum 4, 010331 (2023b).
  • Hayden et al. (2016) P. Hayden, S. Nezami, X.-L. Qi, N. Thomas, M. Walter,  and Z. Yang, Journal of High Energy Physics 2016, 9 (2016).
  • Vasseur et al. (2019) R. Vasseur, A. C. Potter, Y.-Z. You,  and A. W. W. Ludwig, Phys. Rev. B 100, 134203 (2019).
  • Zhou and Nahum (2019) T. Zhou and A. Nahum, Phys. Rev. B 99, 174205 (2019).
  • Nahum and Joerg Wiese (2023) A. Nahum and K. Joerg Wiese, arXiv e-prints , arXiv:2303.07848 (2023)arXiv:2303.07848 [cond-mat.stat-mech] .
  • Li et al. (2021b) Y. Li, X. Chen, A. W. W. Ludwig,  and M. P. A. Fisher, Phys. Rev. B 104, 104305 (2021b).
  • Gurarie (1993) V. Gurarie, Nuclear Physics B 410, 535 (1993).
  • Gurarie and Ludwig (2005) V. Gurarie and A. W. W. Ludwig, in From Fields to Strings: Circumnavigating Theoretical Physics (WORLD SCIENTIFIC, 2005) pp. 1384–1440; arXiv:hep–th/0409105.
  • Cardy (2013) J. Cardy, Journal of Physics A: Mathematical and Theoretical 46, 494001 (2013).
  • Zabalo et al. (2020) A. Zabalo, M. J. Gullans, J. H. Wilson, S. Gopalakrishnan, D. A. Huse,  and J. H. Pixley, Phys. Rev. B 101, 060301 (2020).
  • Sierant et al. (2022b) P. Sierant, M. Schirò, M. Lewenstein,  and X. Turkeshi, Phys. Rev. B 106, 214316 (2022b).
  • Block et al. (2022) M. Block, Y. Bao, S. Choi, E. Altman,  and N. Y. Yao, Phys. Rev. Lett. 128, 010604 (2022).
  • Zabalo et al. (2022) A. Zabalo, M. J. Gullans, J. H. Wilson, R. Vasseur, A. W. W. Ludwig, S. Gopalakrishnan, D. A. Huse,  and J. H. Pixley, Phys. Rev. Lett. 128, 050602 (2022).
  • Calabrese and Cardy (2009) P. Calabrese and J. Cardy, Journal of Physics A: Mathematical and Theoretical 42, 504005 (2009).
  • Cardy (1984a) J. L. Cardy, Nuclear Physics B 240, 514 (1984a).
  • Cardy (1989) J. L. Cardy, Nuclear Physics B 324, 581 (1989).
  • Bertini et al. (2019) B. Bertini, P. Kos,  and T. c. v. Prosen, Phys. Rev. Lett. 123, 210601 (2019).
  • Gopalakrishnan and Lamacraft (2019) S. Gopalakrishnan and A. Lamacraft, Phys. Rev. B 100, 064309 (2019).
  • Piroli et al. (2020) L. Piroli, B. Bertini, J. I. Cirac,  and T. c. v. Prosen, Phys. Rev. B 101, 094304 (2020).
  • Claeys et al. (2022) P. W. Claeys, M. Henry, J. Vicary,  and A. Lamacraft, Phys. Rev. Res. 4, 043212 (2022).
  • Affleck (1986) I. Affleck, Phys. Rev. Lett. 56, 746 (1986).
  • Blöte et al. (1986) H. W. J. Blöte, J. L. Cardy,  and M. P. Nightingale, Phys. Rev. Lett. 56, 742 (1986).
  • Affleck (1988) I. Affleck, in Finite-Size Scaling, Current Physics–Sources and Comments, Vol. 2, edited by J. L. CARDY (Elsevier, 1988) pp. 347–349.
  • Francesco et al. (2012) P. Francesco, P. Mathieu,  and D. Sénéchal, Conformal field theory (Springer Science & Business Media, 2012).
  • Note (1) Because the bulk is critical, any boundary condition implemented at the microscopic lattice scale of the circuit will at long scales result in a scale-invariant and conformally invariant boundary condition.
  • Cardy (1984b) J. L. Cardy, Journal of Physics A: Mathematical and General 17, L385 (1984b).
  • Nielsen and Chuang (2010) M. A. Nielsen and I. L. Chuang, Quantum computation and quantum information (Cambridge university press, 2010).
  • Weinstein et al. (2022) Z. Weinstein, Y. Bao,  and E. Altman, Phys. Rev. Lett. 129, 080501 (2022).
  • Li and Claassen (2023) Y. Li and M. Claassen, arXiv preprint arXiv:2303.08152  (2023).
  • Note (2) For both of these interpretations of the boundary conditions f𝑓fitalic_f and a𝑎aitalic_a, as (simple product) initial state and final state containing the physical qubits, respectively, see Fig. 2 (a), and the corresponding text of Ref. \rev@citealpnumPhysRevB.104.104305.
  • Note (3) paragraph below Eq. 2 .
  • Note (4) see, e.g., Ref. \rev@citealpnumPhysRevB.104.104305.
  • Note (5) see Ref. \rev@citealpnumPhysRevB.104.104305, Appendix C, and Ref. \rev@citealpnumPhysRevA.71.042306.
  • Popp et al. (2005) M. Popp, F. Verstraete, M. A. Martín-Delgado,  and J. I. Cirac, Phys. Rev. A 71, 042306 (2005).
  • Lin et al. (2023) C.-J. Lin, W. Ye, Y. Zou, S. Sang,  and T. H. Hsieh, Quantum 7, 910 (2023).
  • Gullans and Huse (2020b) M. J. Gullans and D. A. Huse, Phys. Rev. Lett. 125, 070606 (2020b).
  • Nahum and Skinner (2020) A. Nahum and B. Skinner, Phys. Rev. Res. 2, 023288 (2020).
  • Lang and Büchler (2020) N. Lang and H. P. Büchler, Phys. Rev. B 102, 094204 (2020).
  • Ippoliti et al. (2021) M. Ippoliti, M. J. Gullans, S. Gopalakrishnan, D. A. Huse,  and V. Khemani, Phys. Rev. X 11, 011030 (2021).
  • Lavasani et al. (2021b) A. Lavasani, Y. Alavirad,  and M. Barkeshli, Nature Physics 17, 342 (2021b).
  • Sang and Hsieh (2021) S. Sang and T. H. Hsieh, Physical Review Research 3, 023200 (2021).
  • Sang et al. (2021) S. Sang, Y. Li, T. Zhou, X. Chen, T. H. Hsieh,  and M. P. Fisher, PRX Quantum 2, 030313 (2021).
  • Sriram et al. (2022) A. Sriram, T. Rakovszky, V. Khemani,  and M. Ippoliti, arXiv preprint arXiv:2207.07096  (2022).
  • Newman (1994) C. M. Newman, Probability and phase transition , 247 (1994).
  • Aaronson and Gottesman (2004) S. Aaronson and D. Gottesman, Phys. Rev. A 70, 052328 (2004).
  • Gottesman (1998) D. Gottesman, arXiv preprint quant-ph/9807006  (1998).
  • Nahum et al. (2017) A. Nahum, J. Ruhman, S. Vijay,  and J. Haah, Phys. Rev. X 7, 031016 (2017).
  • Krastanov et al. (2022) S. Krastanov, P. Viswanathan, J. Lapeyre, C. Zhao, Rabqubit, ShuGe-MIT,  and gsommers, “Krastanov/quantumclifford.jl: v0.6.3,”  (2022).
  • Jacobsen and Saleur (2008) J. L. Jacobsen and H. Saleur, Nuclear Physics B 788, 137 (2008).
  • Barends et al. (2014) R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E. Jeffrey, T. C. White, J. Mutus, A. G. Fowler, B. Campbell, et al., Nature 508, 500 (2014).
  • Pordes et al. (2007) R. Pordes, D. Petravick, B. Kramer, D. Olson, M. Livny, A. Roy, P. Avery, K. Blackburn, T. Wenaus, F. Würthwein, I. Foster, R. Gardner, M. Wilde, A. Blatecky, J. McGee,  and R. Quick, in J. Phys. Conf. Ser., 78, Vol. 78 (2007) p. 012057.
  • Sfiligoi et al. (2009) I. Sfiligoi, D. C. Bradley, B. Holzman, P. Mhashilkar, S. Padhi,  and F. Wurthwein, in 2009 WRI World Congress on Computer Science and Information Engineering, 2, Vol. 2 (2009) pp. 428–432.