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

Entanglement negativity at the critical point of measurement-driven transition

Bowen Shi Department of Physics, The Ohio State University, Columbus, OH 43210, USA Department of Physics, University of California at San Diego, La Jolla, CA 92093, USA    Xin Dai Department of Physics, The Ohio State University, Columbus, OH 43210, USA Computational Science Initiative, Brookhaven National Laboratory, Upton, NY 11973, USA    Yuan-Ming Lu Department of Physics, The Ohio State University, Columbus, OH 43210, USA
Abstract

We study the entanglement behavior of a random unitary circuit punctuated by projective measurements at the measurement-driven phase transition in one spatial dimension. We numerically study the logarithmic entanglement negativity of two disjoint intervals and find that it scales as a power of the cross-ratio. We investigate two systems: (1) Clifford circuits with projective measurements, and (2) Haar random local unitary circuit with projective measurements. Remarkably, we identify a power-law behavior of entanglement negativity at the critical point. Previous results of entanglement entropy and mutual information point to an emergent conformal invariance of the measurement-driven transition. Our result suggests that the critical behavior of the measurement-driven transition is distinct from the ground state behavior of any unitary conformal field theory.

I Introduction and main results

Entanglement is a central concept in quantum physics, and it enables some of the most important applications of quantum mechanics such as quantum teleportation and quantum computation. For bipartite pure states, the entanglement entropy measures the entanglement of subsystem AA with its complement. In recent years, the entanglement entropy has proven to be an insightful tool to characterize quantum many-body phases[1, 2, 3, 4, 5].

The entanglement entropy, however, is not a complete characterization of quantum entanglement[6, 7, 8]. While entanglement entropy and associated mutual information provide a good characterization of entanglement in a pure quantum state, they measure the total amount of correlation rather than the quantum entanglement between two regions in a mixed quantum state[9, 10]. How to quantify the quantum entanglement for bipartite mixed states (or equivalently tripartite pure states)? One such calculable measure is known as the entanglement negativity[11, 6]. Recently entanglement negativity as well as other mixed states entanglement measures have been applied to enrich our understanding of various quantum many-body systems[12, 13, 14, 15, 16, 17, 18, 19].

Recently, there are significant interests in the entanglement properties of a 1D dynamical quantum system, driven by random local unitary circuits and random local measurements with a certain rate 0<p<10<p<1 [20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33]. In such random unitary circuits with measurements, the unitary evolution tends to increase the entanglement entropy [34] whereas the local measurements tend to disentangle the system. This leads to a volume-law entangled phase at a low rate of measurement p<pcp<p_{c} and an area-law entangled phase at a high rate of measurement p>pcp>p_{c} [20, 21, 22]. After that, a number of works have studied various properties of the two phases and the critical point that separates them [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33]. Most of the previous works focused on the entanglement entropy and mutual information between different regions in the system.

In this work, our goal is to characterize the quantum entanglement in the 1D random unitary circuits with measurements by numerically studying the entanglement negativity between two disjoint intervals. In particular, we focus on the critical behavior of the entanglement negativity at the measurement-driven phase transition at p=pcp=p_{c}. We investigate both the Clifford random unitary circuit and the Haar random unitary circuit, punctuated by single-site projective measurements at a rate of pp. We numerically identify a simple scaling law of the entanglement negativity right at the critical point. For two small intervals separated by a distance rr, the negativity scales as r2Λr^{-2\Lambda}, where Λ3\Lambda\approx 3 for both Clifford and Haar circuit models (within the error bar). Alternatively, the scaling law can be written as ηΛ\eta^{\Lambda} for η1\eta\ll 1, where η\eta is the cross-ratio defined by (6) for the pair of intervals.

It is insightful to compare the power-law scaling of negativity studied in this work to previously obtained power-law scaling of mutual information at the same measurement-driven phase transition [23]. The mutual information (2) at the critical point has been shown to scale in a power-law fashion IA,BηΔI_{A,B}\propto\eta^{\Delta} with the cross-ratio η\eta for η1\eta\ll 1, where Δ2\Delta\approx 2. In other words, the mutual information for two small disjoint intervals decays as r2Δr^{-2\Delta} with respect to the distance rr between the two intervals; see also Ref. [21]. Δ<Λ\Delta<\Lambda indicates that the entanglement negativity decays faster than the mutual information as we increase the distance between the two intervals. For a mixed quantum state, the bipartite mutual information IA,BI_{A,B} measures the total correlation between the two regions[9, 10], whereas the negativity as an entanglement monotone measures the quantum entanglement between them[35, 11, 6]. Therefore our results indicate as the distance between two regions increases, that the quantum entanglement decays faster than the total correlation between them.

The most important implication of the power-law negativity scaling is on the nature of the measurement-driven phase transition. Many previous results, such as the scalings of entanglement entropy and mutual information, point to an emergent conformal invariance at this dynamical critical point[23, 33, 25]. While the precise nature of this emergent conformal symmetry is under debate, it remained possible that the entanglement property of the late time quantum states being similar to that in the continuous phase transitions in equilibrium, at least phenomenologically. In other words, the long-wavelength physics of the measurement-driven entanglement transition may be captured by that of the ground state of some conformal field theory (CFT). While the power-law mutual information is consistent with a unitary CFT[36, 37, 37, 38], it is known that the entanglement negativity of two disjoint intervals must decay faster than any power law w.r.t. a small cross-ratio η1\eta\ll 1 in a unitary CFT ground state[12, 13, 39]. Therefore our observation on a power-law scaling of entanglement negativity shows strong evidence that the critical behavior at the measurement-driven entanglement transition is distinct from the ground state behavior of any unitary CFT. This puts constraints on the nature that conformal symmetry enters this story. This analysis can also provide a constraint on the possible non-unitary CFT description of the critical point[25, 33].

This paper is organized as follows. In Sec. II, we setup the Clifford and Haar random circuit models with projective measurements and introduce the correlation and entanglement measures (mutual information and entanglement negativity) used in this work. In Sec. III, we present our numerical results on the scaling behavior of mutual information and entanglement negativity at the measurement-driven phase transition and discuss their implications. In Sec. IV, we summarize the main conclusions of our work and outlook for the future.

II Setup and background

II.1 The models

We consider a 1D system consisting of LL qubits arranged on a ring with a periodic boundary condition (PBC), as illustrated in Fig. 1. We assume LL is even throughout the paper. The PBC makes it easy to study the dependence of entanglement on the cross-ratio. We study two types of models: (1) the random Clifford circuit with single-site projective measurements, (2) the Haar random unitary circuit model with single-site projective measurements. We fix the initial state to be |ψ(0)=i=1L|0i|\psi(0)\rangle=\otimes_{i=1}^{L}|0\rangle_{i}, i.e. the product state with all spin up in the Pauli ZZ-basis. The details of the two models are described below.

Figure 1: The 1D system consists of LL qubits (sites) arranged on a circle with the periodic boundary condition.

II.1.1 Random Clifford circuits with measurements

The dynamic of the system is given by random Clifford unitary circuits on pairs of nearby qubits and projective measurements of single-site Pauli ZZ operators independently with probability pp for each site at any time step. See Fig. 2 for an illustration of the bricklayer arrangement of the 2-qubit unitary operators and the position of measurements in discretized spacetime.

The advantage of the Clifford circuit model is that it allows an efficient simulation on classical computers [40]. We have simulated the Clifford circuit model for system sizes up to size L=768L=768. Similar to previous studies, our physical quantities are obtained by taking an ensemble average over late time (pure) quantum states. We shall refer to this average as late time average.

random unitaryZZ-measurement
Figure 2: A spacetime diagram for the 1D randum Clifford (Haar) circuit model with projective measurements. The blue blocks are 2-qubit random unitary operators. Each white dot is a single-qubit projective measurement in the ZZ-basis. This type of measurements occurs with a probability pp in every discrete time step.

II.1.2 Haar random unitary circuit with measurements

We also consider a more random evolution on the same physical system, namely the Haar random circuit with measurements. Now each blue block in Fig. 2 refers to a 2-qubit unitary in U(4)U(4) group, selected randomly with respect to the Haar measure of the Lie group U(4)U(4). The measurements are still in the ZZ-basis as before, and these measurements remain independent with a probability pp. The Haar random circuit model is much more computationally expensive, and we have simulated this model up to size L=20L=20.

II.2 Entanglement and correlation measures

In this section, we briefly introduce the measures of entanglement and correlation studied in this work. Among them, entanglement entropy is an entanglement measure of bipartite pure states, mutual information is a measure of the total correlation for a bipartite mixed state [9, 10], and the entanglement negativity is an entanglement measure of bipartite mixed states [6, 7, 8].

II.2.1 Entanglement entropy and mutual information

First we introduce the entanglement entropy and mutual information. The entanglement entropy, or von Neumann entropy, is defined as

S(ρA)=Tr(ρAlog2ρA)S(\rho_{A})=-{\rm Tr}(\rho_{A}\log_{2}\rho_{A}) (1)

where ρA\rho_{A} is the (reduced) density matrix on (sub)system AA. We shall sometimes drop the state label and denote the entropy as SAS_{A}. Suppose the whole system is ABA\cup B. For a pure quantum state of the system, the von Neumann entropy SAS_{A} characterizes the entanglement between AA and BB. In other words, the von Neumann entropy is an entanglement measure for bipartite pure states. For a mixed state on ABA\cup B, on the other hand, the von Neumann entropy SAS_{A} does not characterize the entanglement between AA and BB [6, 7, 8].

The mutual information is defined as

IA,B=SA+SBSAB.I_{A,B}=S_{A}+S_{B}-S_{AB}. (2)

It is a measure of the total amount of correlations between the two subsystems AA and BB, for possibly mixed quantum states [9, 10]. Correlation can be either quantum or classical. For example, a separable state

ρAB=ipiρAiρBi\rho_{AB}=\sum_{i}p_{i}\rho^{i}_{A}\otimes\rho^{i}_{B} (3)

can have classical correlation, but it has no quantum entanglement between AA and BB [6, 7, 8]. Here {ρAi}\{\rho^{i}_{A}\} and {ρBi}\{\rho^{i}_{B}\} are density matrices and {pi}\{p_{i}\} is a probability distribution.

II.2.2 Entanglement negativity

To quantify the entanglement between two subsystems for a possibly mixed quantum state, one may consider a bipartite mixed state entanglement measure. A number of such entanglement measures were introduced and studied; see [7, 8] for a review. These measures must be entanglement monotones, i.e., they satisfy the condition to be non-increasing under local operations and classical communication (LOCC) between the two regions [6, 11]. Furthermore, they vanish for a separable state (3). Therefore these quantities character quantum entanglement rather than the classical correlation for possibly mixed states. Some of these bipartite entanglement measures have very nice theoretical properties. One such example is the squashed entanglement [41], which is additive. However, calculating it requires a minimization process, and therefore it is hard to calculation in general.

Among these proposed bipartite mixed state entanglement measures, the entanglement negativity is a calculable measure [11, 6]. The logarithmic entanglement negativity111An alternative expression, which is called entanglement negativity of state ρ\rho, between regions AA and BB, is defined as 𝒩A|B(ρ)=ρABTA112.\mathcal{N}^{A|B}(\rho)=\frac{\lVert\rho_{AB}^{T_{A}}\rVert_{1}-1}{2}. (4) Throughout the paper, we will stick to the logarithmic entanglement negativity instead of this one. for state ρ\rho, between AA and BB, is defined as

ENA|B(ρ)=log2ρABTA1.E_{N}^{A|B}(\rho)=\log_{2}\lVert\rho_{AB}^{T_{A}}\rVert_{1}. (5)

where TAT_{A} denotes the partial transpose with respect to region AA only, and O1\lVert O\rVert_{1} is the trace norm of operator OO. For a more detailed description, please refer to Appendix A.

In the rest of the paper, we will study the logarithmic entanglement negativity in random unitary circuits with measurements. We shall call it entanglement negativity for short.

III The critical behavior of entanglement negativity

In this section, we discuss our numerical results on the negativity ENA|BE_{N}^{A|B} as well as mutual information IA,BI_{A,B} for the Clifford and Haar circuit models. All physical quantities are calculated as an ensemble average of that on possible late time quantum states. The goal is to understand quantum entanglement properties of the measurement-driven phase transition discovered in Refs.[20, 21, 22, 23]. Previously, numerical results [23, 27] provided excellent evidence that single interval entanglement entropy at the critical point scales with the logarithm of the subsystem size in both models. Here we divide the 1D system into two disjoint intervals AA and BB, and the remaining region AB¯\overline{A\cup B}, and investigate the entanglement and correlation between the intervals AA and BB of the reduced density matrix ρAB\rho_{AB} obtained by tracing out region AB¯\overline{A\cup B}. While the mutual information (2) has been studied previously by Ref.[23], here we focus on the logarithmic entanglement negativity (5), which is known to distinguish quantum entanglement from classical correlations between AA and BB for a mixed state.

To characterize the scaling behavior of the entanglement at the measurement-driven transition, we choose AA and BB to be disjoint arcs of a circle. We use xi[0,2π)x_{i}\in[0,2\pi), i=1,2,3,4i=1,2,3,4 to label the angular positions of the endpoints, as shown in Fig. 3. The cross-ratio is defined as

η=x12x34x13x24,\eta=\frac{x_{12}x_{34}}{x_{13}x_{24}}, (6)

where xijx_{ij} is the chord distance. For periodic boundary condition

xij=Lπsin(πL|xixj|).x_{ij}=\frac{L}{\pi}\sin\left(\frac{\pi}{L}|x_{i}-x_{j}|\right). (7)

We will investigate the scaling of entanglement negativity in the limit of a small cross-ratio η1\eta\ll 1. Previously, the scaling of mutual information on two disjoint intervals has been studied by Ref. [23].

AABBx1x_{1}x2x_{2}x3x_{3}x4x_{4}
Figure 3: A possible choice of disjoint intervals AA and BB. Also illustrated are the labels of the endpoints.

III.1 Determining the critical measurement probability pcp_{c}

AABB|A|=|B|=L/8|A|=|B|=L/8011pcp_{c}volume lawarea lawENA|BE_{N}^{A|B} IA,BI_{A,B}
Figure 4: Schematic phase diagram of the random circuit model with projective measurements, as a function of measurement rate pp. The volume law phase (left) and area law phase (right) are separated by the measurement-driven entanglement transition at pcp_{c}, the critical measurement rate. For relatively small disjoint intervals AA and BB, fixed as two antipodal regions with |A|=|B|=L/8|A|=|B|=L/8 (η0.146\eta\approx 0.146) in this diagram, both the mutual information and the entanglement negativity provide a sharp feature of the critical point. Both IA,BI_{A,B} and ENA|BE_{N}^{A|B} are only nonzero nearby the critical point, and they quickly vanish upon entering either the volume law phase or the area law phase.

Previous studies [20, 21, 22, 23] have established the phase diagram of the Clifford and Haar circuit models and identified the measurement-driven phase transition. There is a highly entangled phase, at a small measurement rate p<pcp<p_{c} with volume law entanglement entropy for a single interval, and a disentangled phase at a large measurement rate p>pcp>p_{c} with an area law entanglement entropy; see Fig. 4 for an illustration. One can numerically determine the critical measurement probability pcp_{c} for both models.

Our strategy of determining pcp_{c} is the based on the observation [23] that the mutual information IA,BI_{A,B} has a sharp peak at pcp_{c} and it vanishes quickly with |ppc||p-p_{c}| once entering the volume (area) law phase for a small cross-ratio η\eta. This is illustrated by the blue curve in Fig. 4. We find that the entanglement negativity exhibits a similar peak at pcp_{c}, shown by the orange curve in Fig. 4. Using these sharp features, we can estimate pcp_{c} for the measurement-driven phase transition. The details of data collapse analysis to determine pcp_{c} in the Clifford circuit model are presented in the Appendix B. The value of pcp_{c}, which we shall use in the remaining part of this paper, is:

  • pc0.16p_{c}\approx 0.16 for the Clifford circuit model, consistent with Ref. [23].

  • pc0.26p_{c}\approx 0.26 for the Haar circuit model with L=20L=20, consistent with Ref. [21, 26].

III.2 Scaling of entanglement negativity at pcp_{c}

The main result of this work is a power-law dependence of entanglement negativity on the cross-ratio (6), right at the critical point. We compare this power-law behavior with that of the mutual information; see Fig. 5 for the plot. We vary the size and position of the interval when we collect data. (Namely, we symmetrically adjust the position of the four points on the ring while keeping two intervals of equal length.) In both the Clifford circuit model and the Haar circuit model, we obtain the same power-law behavior for two disjoint intervals AA and BB in Fig. 3:

  • ENA|BηΛE_{N}^{A|B}\propto\eta^{\Lambda}, where Λ3\Lambda\approx 3.

  • IA,BηΔI_{A,B}\propto\eta^{\Delta}, where Δ2\Delta\approx 2. This result agrees with Ref. [23].

More precisely, we find that

  • ΛClifford=3.04±0.08\Lambda_{\textrm{Clifford}}=3.04\pm 0.08 and ΛHaar=2.73±0.28\Lambda_{\textrm{Haar}}=2.73\pm 0.28.

  • ΔClifford=2.16±0.03\Delta_{\textrm{Clifford}}=2.16\pm 0.03 and ΔHaar=1.89±0.26\Delta_{\textrm{Haar}}=1.89\pm 0.26.

Here the error bars come from the standard deviation of the linear fit.

Refer to caption
Refer to caption
Figure 5: Mutual information IA,BI_{A,B} and (logarithmic) entanglement negativity ENA|BE_{N}^{A|B} scales as a power of cross-ratio η\eta at the critical point, for small η\eta. (We take pc=0.16p_{c}=0.16 and L=512L=512 for Clifford and pc=0.26p_{c}=0.26 and L=20L=20 for Haar.) The first figure suggests that the power-law scaling exponent of the mutual information are ΔClifford=2.16±0.03\Delta_{\textrm{Clifford}}=2.16\pm 0.03 and ΔHaar=1.89±0.26\Delta_{\textrm{Haar}}=1.89\pm 0.26. The second figure suggests that the scaling exponent of the (logarithmic) entanglement negativity are ΛClifford=3.04±0.08\Lambda_{\textrm{Clifford}}=3.04\pm 0.08 and ΛHaar=2.73±0.28\Lambda_{\textrm{Haar}}=2.73\pm 0.28.

Both the entanglement negativity and the mutual information scale as a power of the cross-ratio, albeit that the power Λ3\Lambda\approx 3 is different from Δ2\Delta\approx 2. The mutual information characterizes the total amount of correlation between two regions, including both classical correlation and quantum entanglement. On the other hand, the entanglement negativity quantifies the amount of quantum entanglement between two regions. The scaling dimension Λ>Δ\Lambda>\Delta physically means that the quantum entanglement between two disjoint intervals drops faster than the correlation when the distance between the intervals increases. It is physically reasonable to expect ΛΔ\Lambda\geq\Delta, because the amount of quantum entanglement must be bounded by the total amount of correlations.222Depending on the specific entanglement measure in question, the bound may not look simple. However, in the context of the stabilizer states, which is relevant to the random Clifford circuit model with measurement, a simple bound exists: IA,B2ENA|BI_{A,B}\geq 2E_{N}^{A|B}; see Theorem A.2 in Appendix A.

Can these critical phenomena be captured by an effective field theory? As observed in previous studies[23, 33, 25], the power-law behavior (IA,BηΔI_{A,B}\propto\eta^{\Delta}) of the mutual information (also see Ref. [21] for the 0-th Rényi mutual information) suggests an emergent conformal field theory (CFT) description of the critical point. One way this might work is to match the entanglement behavior of the measurement-driven transition to that of some CFT ground-state. For a unitary CFT ground state, the mutual information of two disjoint intervals is known to depend on the full operator content of the theory [36, 37, 37], and calculation must be done case by case. Nevertheless, if the CFT is compact and unitary, the 2nd Renyi mutual information is rigorously shown to be a power of η\eta at small η\eta, and the power is determined knowing the dimension of the lowest non-unit operator in the CFT [38]. Therefore, the scaling of mutual information is analogous to that for a unitary CFT ground state. The power constrains the nature of any CFT that may describe the measurement-driven transition.

The same question arises for entanglement negativity: what does the scaling of negativity imply about the nature of the measurement-drive transition? As a main result of this paper, we numerically observe a power-law dependence of the logarithmic entanglement negativity (5) on the cross-ratio (6) at the critical point (ENA|BηΛE_{N}^{A|B}\propto\eta^{\Lambda}). For unitary CFT ground states, the entanglement negativity is known to decay faster than any power law of cross-ratio for small η\eta [12, 13, 39]. Thus our result suggests that the entanglement behavior at the critical point of the measurement-driven transition is distinct from that of any unitary CFT ground state.

Nonunitary CFTs [42, 43] are natural candidates of the low-energy effective field theory because they obey the conformal symmetry, and the behavior of entanglement negativity may not obey the same rules as the unitary CFTs [12, 13], as discussed in Ref. [44] and the references therein. However, a careful future study is needed to determine the relevance and the nature of possible emergent nonunitary CFT effective description of the critical point in the Clifford and Haar circuit models. The scaling exponent Λ3\Lambda\approx 3, can provide useful information in constraining the nature of this non-unitary CFT.

III.3 Different “Quantumness” of Haar vs. Clifford circuit models

The Clifford circuit model and the Haar circuit model have almost identical scaling exponents within the error bar, for both the mutual information and the entanglement negativity, as shown in Fig. 5. However, we have observed a slight difference in the “quantumness” of these two models, which we describe below.

For a generic mixed quantum states, while mutual information (2) measures the total amount of (both classical and quantum) correlation between two disjoint intervals AA and BB, the entanglement negativity vanishes for any separable states and therefore characterizes the quantum entanglement between the two intervals. Consequently the ratio of the negativity to the mutual information provides a measure of how much correlation comes from quantum entanglement between two regions AA and BB, i.e. the “quantumness” of the state. To compare the Clifford and the Haar circuit models, we define the following quantity:

R(η)(ENA|B(η)IA,B(η))Haar/(ENA|B(η)IA,B(η))CliffordR(\eta)\equiv\left(\frac{E_{N}^{A|B}(\eta)}{I_{A,B}(\eta)}\right)_{\textrm{Haar}}\left/\left(\frac{E_{N}^{A|B}(\eta)}{I_{A,B}(\eta)}\right)_{\textrm{Clifford}}\right. (8)

to feature the excess amount of “quantumness” in the Haar circuit model as compared to Clifford circuit model. η\eta is the cross-ratio defined in (6).

Our numerical data suggests that R(η)R(\eta) depends weakly on the cross-ratio η\eta:

R(η)=R0ηδ,R0=2.94,δ=0.04±0.39.R(\eta)=R_{0}\eta^{\delta},\quad R_{0}=2.94,\,\delta=-0.04\pm 0.39. (9)

R(η)3R(\eta)\approx 3 in the whole range of cross-ratio that we simulate. While this does not imply any difference in the effective field theory description of the two models, this does suggest that there is a larger fraction of quantum entanglement in the Haar circuit model than the Clifford circuit model. Therefore the Haar circuit model appears to be more quantum than the Clifford circuit model at the measurement-driven transition.

IV Summary and discussion

In this work, we have numerically studied the correlation and entanglement between two disjoint intervals at the measurement-driven phase transition in 1D Clifford and Haar random unitary circuits. We focus on the mutual information (2) and the logarithmic entanglement negativity (5). The late time quantum states are generically mixed on the union of the two intervals. Therefore, the mutual information measures the total correlation between the two intervals, whereas the entanglement negativity characterizes the quantum entanglement between the two intervals. We numerically simulate 1D systems with a periodic boundary condition, up to a system size L=768L=768 for the Clifford circuit model and L=20L=20 for the Haar circuit model. We identified a power-law behavior of the (logarithmic) entanglement negativity at small interval sizes. For a pair of disjoint intervals AA and BB,

ENA|BηΛwith Λ3E_{N}^{A|B}\propto\eta^{\Lambda}\quad\textrm{with }\,\Lambda\approx 3 (10)

at η1\eta\ll 1, where η\eta is the cross-ratio for the pair of disjoint interval defined in (6).

It is interesting to compare (10) with recently obtained power-law behavior of the mutual information [23].

IA,BηΔwith Δ2I_{A,B}\propto\eta^{\Delta}\quad\textrm{with }\,\Delta\approx 2 (11)

at η1\eta\ll 1. The fact that the scaling dimension Λ>Δ\Lambda>\Delta suggests that the quantum entanglement between two disjoint intervals decays faster than the correlation as we increase the distance between the intervals. Moreover, we have observed that the Haar circuit model is “more quantum” than the Clifford circuit model, by comparing the ratio (8) of the entanglement negativity to mutual information in the two models.

As the main conclusion of this work, the power-law dependence of negativity on the cross-ratio indicates that the description of the measurement-driven transition in random unitary circuits is distinct from the ground state of any unitary conformal field theory. Previous studies on the scaling of single interval von Neumann and Renyi (entanglement) entropies, as well as the mutual information for two disjoint intervals, point to an emergent conformal invariance of the critical point[33]. However, the precise nature of the CFT describing the phase transition is unclear[23, 33]. Remarkably, unlike the power-law mutual information between two disjoint intervals, in a unitary CFT ground state, the negativity always decays faster than any power law of cross-ratio η\eta for small η\eta [12, 13]. Therefore, the power-law behavior of entanglement negativity provides direct evidence that the entanglement property of the measurement-driven transition is distinct from that of the ground state of any unitary CFT.

Conformal symmetry may, instead, enter the story through statistical mechanical models. In such cases, the resulting property does not necessarily match the ground state properties [45]. Previously, Vasseur et al. [46] argued a vanishing central charge (c=0c=0) for the CFT corresponding to the measurement-driven transition, and indicated that this CFT belongs to a class of non-unitary Logarithmic CFT. This is achieved by using the replica trick and mapping the entanglement entropy in a random unitary circuit with measurements to the change of free energy of a 2D statistical mechanical model w.r.t. twisting the boundary condition[47, 26, 25, 32]. The critical point of the statistical mechanical model belongs to the (Q!)(Q!)-state Potts model in the Q1Q\rightarrow 1 limit, in the universality class of 2D percolation, perturbed by a relevant 2-hull operator. In the current work, inspired by recent numerical evidence of emergent conformal invariance in the mutual information, we assume the critical behavior of the measurement-driven transition is described by a CFT ground state, analogous to continuous phase transitions in equilibrium. The power-law behavior of entanglement negativity hence suggests this CFT cannot be unitary. Currently, it is not clear how to exactly map the entanglement negativity to a 2D statistical mechanics model, employing the idea of Ref. [46]. This is an interesting question to be addressed in the future.

While the power-law negativity in this work provides an extra constraint on the critical theory, many questions remain open for this entanglement phase transition. For example, how to extract other critical exponents of this dynamical critical point? What is the precise nature of the conformal symmetry that emerged at the critical point of the measurement-driven phase transition? How to understand the scaling power that we observed? We leave these questions to future works.

Acknowledgments

We thank Tarun Grover, Matthew Fisher, Brian Skinner, Sagar Vijay, Xueda Wen, Yahya Alavirad, John McGreevy, Yi-Zhuang You for interesting discussions. We are especially grateful to Romain Vasseur for his helpful comment on the nature of the conformal field theory in statistical mechanics models. This work is supported by the National Science Foundation under Grant No. NSF DMR-1653769 (BS, XD, YML), University of California Laboratory Fees Research Program, grant LFR-20-653926 (BS), and the Simons Collaboration on Ultra-Quantum Matter, grant 651440 from the Simons Foundation (BS).

Note: During the preparation of this paper, we became aware of an independent work by Shengqi Sang, Yaodong Li, Tianci Zhou, Xiao Chen, Timothy H. Hsieh, and Matthew P. A. Fisher, who also studied entanglement negativity at the same measurement-driven transition in random unitary circuits (Ref. [48]).

Appendix A Entanglement of stabilizer states

In this appendix, we provide some details of the entanglement of stabilizer states. In Section A.1, we provide a brief review of the stabilizer states and set up the notation for the remaining discussion. In Section A.2, we review the calculation of entanglement entropy (bipartite pure state entanglement measure) of the stabilizer states. In Section A.3, A.4 and A.5, we describe a method to calculation entanglement negativity (a bipartite mixed state entanglement measure) of stabilizer states. While this appendix is relatively self-contained, we mention that various statements in this appendix may be seen from a useful alternative viewpoint on the multipartite entanglement of stabilizer states [49].

As a reminder, Clifford unitary operators and projective measurements in the ZZ-basis, when applied to a stabilizer state, corresponds to an update of the stabilizer generators. The state after each step of evolution is again a stabilizer state. Therefore, the states we consider in the random Clifford circuit model with measurements are stabilizer states. For a review of how to update the stabilizers for Clifford unitary evolution and projective measurement, see [40, 50] and appendices of [23] and [34].

A.1 Stabilizer states

We will consider a quantum system built up from a set of qubits. The number of qubit of the whole system is LL. In other words, the total Hilbert space is

=2L,2=span{|0,|1}.\mathcal{H}=\mathcal{H}_{2}^{\otimes L},\quad\mathcal{H}_{2}=span\{|0\rangle,|1\rangle\}. (12)

Below, we define three concepts: stabilizer group, stabilizer generators, and stabilizer states.

Definition A.1.

(stabilizer group) The stabilizer group 𝒢\mathcal{G} is defined to be the Abelian group generated by the following operators {g1,g2,,gL}\{g_{1},g_{2},\cdots,g_{L}\} acting on {\mathcal{H}}. These operators, which we shall call stabilizer generators, satisfy:

  1. 1.

    Each gig_{i} is a product of Pauli operators. In other words, it is a product of operators acting on individual sites, and on each site, the operator (up to a phase factor) can be chosen from the set {I,X,Y,Z}\{I,X,Y,Z\}. Here II is the identity, XX, YY, ZZ are single-qubit Pauli operators.

  2. 2.

    gi=gig_{i}^{\dagger}=g_{i} and gigj=gjgig_{i}g_{j}=g_{j}g_{i}, i,j\forall i,j.

  3. 3.

    The set of operators are independent, i.e.

    gijigjsj,sj=0,1g_{i}\neq\prod_{j\neq i}g_{j}^{s_{j}},\quad s_{j}=0,1 (13)
Remark.

The definition implies gi2=1g_{i}^{2}=1. It follows from the definition that |𝒢|=2L|\mathcal{G}|=2^{L}, where |𝒢||\mathcal{G}| is the number of group element. Due to the independence condition, the stabilizer generators can independently take ±1\pm 1. For each choice, there is a unique quantum state (up to the overall phase factor) that has these as the eigenvalues.

Definition A.2.

(stabilizer Hamiltonian) We define the following Hamiltonian as the stabilizer Hamiltonian of 𝒢\mathcal{G}:

H=j=1Lgj,H=-\sum_{j=1}^{L}\,g_{j}, (14)

where {g1,g2,,gL}\{g_{1},g_{2},\cdots,g_{L}\} is the set of stabilizer generators.

Definition A.3.

(stabilizer state) The stabilizer state of 𝒢\mathcal{G} is the unique quantum state (up to an overall phase factor) |ψ|\psi\rangle that satisfies

h|ψ=|ψ,h𝒢.h|\psi\rangle=|\psi\rangle,\quad\forall h\in\mathcal{G}. (15)
Remark.

The stabilizer state is the unique ground state of the stabilizer Hamiltonian (14).

We define the following projector:

P1|𝒢|h𝒢hP=1|𝒢|j=1L(1+gj).P\equiv\frac{1}{|\mathcal{G}|}\sum_{h\in\mathcal{G}}h\quad\Leftrightarrow\quad P=\frac{1}{|\mathcal{G}|}\prod_{j=1}^{L}(1+g_{j}). (16)

It satisfies

P\displaystyle P =\displaystyle= P\displaystyle P^{\dagger} (17)
P2\displaystyle P^{2} =\displaystyle= P.\displaystyle P. (18)

In fact, PP is the projector to the ground subspace of the stabilizer Hamiltonian and therefore

P=|ψψ|.P=|\psi\rangle\langle\psi|. (19)

In this way, we get a neat formula for the ground state density matrix.

Let us consider a subsystem AA of the system and let the system be ABAB. Let LAL_{A} be the number of qubits in AA. Let 𝒢A\mathcal{G}_{A} be the subgroup of stabilizers supported on AA. See the formal definition below.

Definition A.4.
𝒢A{hA|hA1B𝒢}.\mathcal{G}_{A}\equiv\{h_{A}\,|h_{A}\otimes 1_{B}\in\mathcal{G}\}. (20)

It is easy to say 𝒢A\mathcal{G}_{A} is isomorphic to a subgroup of 𝒢\mathcal{G}.

The reduced density matrix of the stabilizer state on subsystem AA, ρATrB|ψψ|\rho_{A}\equiv{\rm Tr}_{B}|\psi\rangle\langle\psi|, can be written as

ρA=|𝒢A|2LAPA,PA1|𝒢A|hA𝒢AhA\rho_{A}=\frac{|\mathcal{G}_{A}|}{2^{L_{A}}}P_{A},\quad P_{A}\equiv\frac{1}{|\mathcal{G}_{A}|}\sum_{h_{A}\in\mathcal{G}_{A}}h_{A} (21)

or simply

ρA=12LAhA𝒢AhA.\rho_{A}=\frac{1}{2^{L_{A}}}\sum_{h_{A}\in\mathcal{G}_{A}}h_{A}. (22)

One easily verifies that PAP_{A} in (21) is a projector, and therefore

ρA2=|𝒢A|2LAρA.\rho_{A}^{2}=\frac{|\mathcal{G}_{A}|}{2^{L_{A}}}\rho_{A}. (23)

A density matrix satisfying (23) must have a flat entanglement spectrum. In other words, all the nonzero eigenvalues of ρA\rho_{A} are the same: λ=|𝒢A|/2LA\lambda={|\mathcal{G}_{A}|}/{2^{L_{A}}}. Furthermore, |𝒢A|=2mA|\mathcal{G}_{A}|=2^{m_{A}} for some nonnegative integer mAm_{A}.

A.2 Entanglement entropy of stabilizer states

With the knowledge of the stabilizer states discussed above, it is straightforward to calculation its entanglement entropy. For a stabilizer state ρA\rho_{A}, defined in (21), we have

S(ρA)=Sα(ρA)=(LAmA),α[0,1)(1,)S(\rho_{A})=S_{\alpha}(\rho_{A})=(L_{A}-m_{A}),\quad\alpha\in[0,1)\cup(1,\infty) (24)

Here S(ρA)=Tr(ρAlog2ρA)S(\rho_{A})=-{\rm Tr}(\rho_{A}\log_{2}\rho_{A}) is the von Neumann entropy and SαS_{\alpha} is the Renyi entropy of order α\alpha, defined as

Sα(ρA)=11αlog2[Tr(ρA)α].S_{\alpha}(\rho_{A})=\frac{1}{1-\alpha}\log_{2}[{\rm Tr}\,(\rho_{A})^{\alpha}]. (25)

Note that Renyi entropy is not defined for α=1\alpha=1. Nevertheless, Renyi entropy at the limit α1\alpha\to 1 (in both 1+1^{+} and 11^{-} direction) gives us the von Neumann entropy. As an aside, we can infer from (24) that mALAm_{A}\leq L_{A}.

A practically useful method is to write the binary vectors of a known set of stabilizer generators {g1,,gL}\{g_{1},\cdots,g_{L}\} as a L×2LL\times 2L matrix over F2F_{2}:

M=(MA|MB),M=(M_{A}|M_{B}), (26)

where MAM_{A} is a L×2LAL\times 2L_{A} matrix and MBM_{B} is a L×2LBL\times 2L_{B} matrix; we have partitioned the systems into AA and BB of length LAL_{A} and LB=LLAL_{B}=L-L_{A}. Here each stabilizer generator is mapped to a row of length 2L2L. (Explicitly, we shall choose the binary representations I=(00)I=(00), X=(10)X=(10), Z=(01)Z=(01), Y=(11)Y=(11).)

We have the formula:

mA=LABRank[2](MB),m_{A}=L_{AB}-{\rm Rank}_{[2]}(M_{B}), (27)

where the rank Rank[2](MB){\rm Rank}_{[2]}(M_{B}) is defined over field F2F_{2}.

It is easy to see, the von Neumann entropy of a stabilizer state, ρA\rho_{A} of (21), is

S(ρA)=Rank[2](MA)LA.S(\rho_{A})={\rm Rank}_{[2]}(M_{A})-L_{A}. (28)
Example A.1.

Let L=2L=2 and the set of stabilizer generators be {XX,ZZ}\{XX,ZZ\}. Then the matrix MM reads

M=(10100101).M=\left(\begin{array}[]{cccc}1&0&1&0\\ 0&1&0&1\end{array}\right). (29)

Let LA=1L_{A}=1, then

MA=(1001)Rank[2](MA)=2.M_{A}=\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right)\quad\Rightarrow\quad{\rm Rank}_{[2]}(M_{A})=2. (30)

According to (28), we have S(ρA)=1S(\rho_{A})=1.

A.3 Entanglement negativity basics

Entanglement negativity is a bipartite mixed state entanglement measure. While it goes back to (Renyi α=1/2\alpha=1/2) entanglement entropy for a pure state, for a mixed state it is in general different from any linear combinition of entanglement entropies. It characterizes the quantum entanglement between two regions AA and BB for a possibly mixed state ρAB\rho_{AB}, rather than the correlations between AA and BB. For example, we may consider a separable state

ρAB=ipiρAiρBi\rho_{AB}=\sum_{i}p_{i}\,\rho^{i}_{A}\otimes\rho_{B}^{i} (31)

where {pi}\{p_{i}\} is a probability distribution and ρAi\rho^{i}_{A}, ρBi\rho^{i}_{B} are density matrices. In general, a separable state can have nontrivial correlation between the two subsystems because, it is possible that the mutual information

IA,B(ρAB)S(ρA)+S(ρB)S(ρAB)>0.I_{A,B}(\rho_{AB})\equiv S(\rho_{A})+S(\rho_{B})-S(\rho_{AB})>0. (32)

However, for a separable state, the correlation is classical. This correlation can be generated from local operation and classical communication (LOCC). Therefore, it cannot be used as a resource for quantum teleportation.

As a quantity non-increasing under LOCC [11], (logarithmic) entanglement negativity can characterize the quantum entanglement between AA and BB, for a mixed state ρAB\rho_{AB}.

Definition A.5.

(logarithmic entanglement negativity) For ρAB\rho_{AB}, let us define the logarithmic entanglement negativity as

ENA|B(ρAB)=log2ρABTA1.E_{N}^{A|B}(\rho_{AB})=\log_{2}\lVert\rho_{AB}^{T_{A}}\rVert_{1}. (33)

Below, we explain what is TAT_{A} and the trace norm A1\lVert A\rVert_{1}.

To define TAT_{A}, we need to specify an orthonormal basis of subsystem AA, and the operation TAT_{A} will (in general) be different if we pick a different basis. (Nevertheless, one can show that different TAT_{A} will result in the same trace norm ρABTA1\lVert\rho_{AB}^{T_{A}}\rVert_{1}. For this reason, we do not need to care about the basis choice.

Let us pick an orthonormal basis {|iA}\{|i_{A}\rangle\} of A\mathcal{H}_{A} for TAT_{A} and for convenience, we also pick an orthonormal basis {|jB}\{|j_{B}\rangle\} for B\mathcal{H}_{B}. Then, TAT_{A} can be defined as the linear transformation on the operators acting on AB\mathcal{H}_{AB} that satisfies

(|iA,jBiA,jB|)TA=|iA,jBiA,jB|.(|i_{A},j_{B}\rangle\langle i^{\prime}_{A},j^{\prime}_{B}|)^{T_{A}}=|i^{\prime}_{A},j_{B}\rangle\langle i_{A},j^{\prime}_{B}|. (34)

The trace norm A1\lVert A\rVert_{1} is defined as

A1TrAA.\lVert A\rVert_{1}\equiv{\rm Tr}\sqrt{A^{\dagger}A}. (35)

When AA is a Hermitian operator, A1\lVert A\rVert_{1} equals to sum of the absolute values of the eigenvalues of AA. Note that ρABTA\rho_{AB}^{T_{A}} is a Hermitian operator when ρAB\rho_{AB} is a density matrix.

Remark.

If ρAB=|ψABψAB|\rho_{AB}=|\psi_{AB}\rangle\langle\psi_{AB}| is a pure state density matrix, then

ENA|B(ρAB)=S12(ρAB).E_{N}^{A|B}(\rho_{AB})=S_{\frac{1}{2}}(\rho_{AB}). (36)

Entanglement negativity is more interesting if we consider a tripartite pure state |ψABC|\psi_{ABC}\rangle or consider a mixed state ρAB\rho_{AB}.

A.4 Entanglement negativity of stabilizer states

Let |ψABC|\psi_{ABC}\rangle be a stabilizer state on a tripartite system ABCABC. We trace out the subsystem CC and get

ρAB=|𝒢AB|2LABPAB,PAB1|𝒢AB|hAB𝒢ABhAB.\rho_{AB}=\frac{|\mathcal{G}_{AB}|}{2^{L_{AB}}}P_{AB},\quad P_{AB}\equiv\frac{1}{|\mathcal{G}_{AB}|}\sum_{h_{AB}\in\mathcal{G}_{AB}}h_{AB}. (37)

This fact has been discussed around (21).

Let us denote the set of stabilizer generators of the subgroup 𝒢AB\mathcal{G}_{AB} as {hABi}i=1mAB\{h_{AB}^{i}\}_{i=1}^{m_{AB}}, where the integer mABm_{AB} is the number of stabilizer generators of 𝒢AB\mathcal{G}_{AB}. Then, a question is how to calculate ENA|B(ρAB)E_{N}^{A|B}({\rho_{AB}}) efficiently given these stabilizer generators.

We find that it is useful to define a mAB×mABm_{AB}\times m_{AB} symmetric matrix JJ over field F2F_{2}:

Jij={1{hAi,hAj}=0,0otherwise,J_{ij}=\left\{\begin{array}[]{ll}1&\{h^{i}_{A},h^{j}_{A}\}=0,\\ 0&\textrm{otherwise,}\end{array}\right. (38)

where we have factorized the stabilizer generators as

hABi=hAihBi.h_{AB}^{i}=h_{A}^{i}\otimes h_{B}^{i}. (39)

Note that, Jii=0J_{ii}=0 for all ii. Therefore, JJ may also be treated as a skew-symmetric matrix over F2F_{2}.

Theorem A.1.

For a stabilizer density matrix ρAB\rho_{AB} in (37),

ENA|B(ρAB)=12Rank[2](J),E_{N}^{A|B}(\rho_{AB})=\frac{1}{2}{\rm Rank}_{[2]}(J), (40)

where Rank[2](J){\rm Rank}_{[2]}(J) is the rank of JJ over field F2F_{2}.

Theorem A.1 is useful in the calculation of the entanglement negativity of stabilizer states. See Sec. A.5 below for the proof.

Theorem A.2.

For a stabilizer density matrix ρAB\rho_{AB} in (37),

2ENA|B(ρAB)IA,B(ρAB).2E_{N}^{A|B}(\rho_{AB})\leq I_{A,B}(\rho_{AB}). (41)
Remark.

The same bound does not generalize to arbitrary quantum states. A simple way to see this is to observe that we can find a pure state |φAB|\varphi_{AB}\rangle on a 2-qubit system such that S12(ρA)λS(ρA)S_{\frac{1}{2}}(\rho_{A})\geq\lambda S(\rho_{A}) for any real number λ\lambda. Here ρA\rho_{A} is the reduced density matrix of |φAB|\varphi_{AB}\rangle.

Proof.

We only need to show

Rank[2](J)mABmAmB.{\rm Rank}_{[2]}(J)\leq m_{AB}-m_{A}-m_{B}. (42)

This is because we can rewrite the left-hand side of (41) using Theorem A.1 and rewrite the right-hand side using (24).

For the calculation of Rank[2](J){\rm Rank}_{[2]}(J), we have the freedom to choose the set of stabilizer generators of 𝒢AB\mathcal{G}_{AB}. Let us choose the following set

{hAi1B}i=1mA{1AhBj}j=1mB{lAkrBk}k=1mABmAmB,\{h_{A}^{i}\otimes 1_{B}\}_{i=1}^{m_{A}}\cup\{1_{A}\otimes h_{B}^{j}\}_{j=1}^{m_{B}}\cup\{l_{A}^{k}\otimes r_{B}^{k}\}_{k=1}^{m_{AB}-m_{A}-m_{B}}, (43)

where {hAi}i=1mA\{h_{A}^{i}\}_{i=1}^{m_{A}} is the set of generators of 𝒢A\mathcal{G}_{A} and {hBj}j=1mB\{h_{B}^{j}\}_{j=1}^{m_{B}} is the set of generators of 𝒢B\mathcal{G}_{B}. We have factorized the stabilizer generators of 𝒢AB\mathcal{G}_{AB} into a product on AA and BB.

In the basis (43), the matrix JJ takes a simple form

J=(0)ma(0)mbJ,J=\left(\begin{array}[]{c}0\end{array}\right)^{\oplus m_{a}}\oplus\left(\begin{array}[]{c}0\end{array}\right)^{\oplus m_{b}}\oplus J^{\prime}, (44)

where JJ^{\prime} is a symmetric (mABmAmB)×(mABmAmB)(m_{AB}-m_{A}-m_{B})\times(m_{AB}-m_{A}-m_{B}) matrix over field F2F_{2}, defined according to

Jkk={1{lAk,lAk}=0,0otherwise.J^{\prime}_{kk^{\prime}}=\left\{\begin{array}[]{ll}1&\{l^{k}_{A},l^{k^{\prime}}_{A}\}=0,\\ 0&\textrm{otherwise.}\end{array}\right. (45)

It is obvious from (44) that Rank[2](J)mABmAmB{\rm Rank}_{[2]}(J)\leq m_{AB}-m_{A}-m_{B}. This completes the proof. ∎

Below are two simple examples that illustrate the calculation.

Example A.2.

Let LA=LB=1L_{A}=L_{B}=1 and the set of stabilizer generators of 𝒢AB\mathcal{G}_{AB} be {XX,ZZ}\{XX,ZZ\}. Then the matrix JJ is 2×22\times 2:

J=(0110)Rank[2](J)=2.J=\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right)\quad\Rightarrow\quad{\rm Rank}_{[2]}(J)=2. (46)

Therefore, according to Theorem A.1, ENA|B(ρAB)=1E_{N}^{A|B}(\rho_{AB})=1. This result makes sense because the stabilizer state for this case is a Bell state

|ψAB=12(|00+|11),|\psi_{AB}\rangle=\frac{1}{\sqrt{2}}(|00\rangle+|11\rangle), (47)

written in the ZZ-basis.

Example A.3.

Let LA=LB=LC=1L_{A}=L_{B}=L_{C}=1 and the set of stabilizer generators of 𝒢\mathcal{G} be {ZZI,IZZ,XXX}\{ZZI,IZZ,XXX\}. Then the subgroup 𝒢AB\mathcal{G}_{AB} has a unique generator ZZZZ, and therefore the matrix JJ is 1×11\times 1:

J=(0)Rank[2](J)=0.J=\left(\begin{array}[]{c}0\end{array}\right)\quad\Rightarrow\quad{\rm Rank}_{[2]}(J)=0. (48)

Therefore, according to Theorem A.1, ENA|B(ρAB)=0E_{N}^{A|B}(\rho_{AB})=0. This result makes sense because the stabilizer state for this case is the GHZ state

|ψABC=12(|000+|111),|\psi_{ABC}\rangle=\frac{1}{\sqrt{2}}(|000\rangle+|111\rangle), (49)

written in the ZZ-basis.

A.5 The proof of Theorem A.1

The following three lemmas directly lead to the proof of Theorem A.1.

Lemma A.3.

Rank[2](J){\rm Rank}_{[2]}(J) is invariant under the change of the generators of 𝒢AB\mathcal{G}_{AB}.

Proof.

Any change of generators can be done by a sequence of (1) permutations of generators (2) multiplying one generator to another. These operations induce an operation on the matrix JJ:

  1. 1.

    Switch two rows aba\leftrightarrow b and then switch two columns aba\leftrightarrow b. (Here, aba\neq b.)

  2. 2.

    Add the aa-th row to the bb-th row and then add the aa-th column to the bb-th column over the field F2F_{2}. (Here, aba\neq b.)

It is easy to see that these two operations do not change Rank[2](J){\rm Rank}_{[2]}(J). This completes the proof. ∎

Lemma A.4.

It is possible to choose a “standard basis” of stabilizer generators of 𝒢AB\mathcal{G}_{AB}:

i=1ma{aABi,bABi}s=1mc{cABs},\cup_{i=1}^{m_{a}}\{a^{i}_{AB},b^{i}_{AB}\}\cup_{s=1}^{m_{c}}\{c^{s}_{AB}\}, (50)

such that

  1. 1.

    The nonnegative integers mam_{a} and mcm_{c} satisfy

    2ma+mc=mAB.2m_{a}+m_{c}=m_{AB}. (51)
  2. 2.

    When we write

    aABi\displaystyle a^{i}_{AB} =aAiaBi,\displaystyle=a^{i}_{A}\otimes a^{i}_{B}, (52)
    bABj\displaystyle b^{j}_{AB} =bAjbBj,\displaystyle=b^{j}_{A}\otimes b^{j}_{B},
    cABs\displaystyle c^{s}_{AB} =cAscBs,\displaystyle=c^{s}_{A}\otimes c^{s}_{B},

    we have

    {aAi,bAi}=0,\displaystyle\,\{a_{A}^{i},b_{A}^{i}\}=0, i,\displaystyle\quad\forall i, (53)
    [aAi,bAj]=0,\displaystyle\,[a_{A}^{i},b_{A}^{j}]=0, ij,\displaystyle\quad i\neq j,
    [aAi,aAj]=[bAi,bAj]=0,\displaystyle\,[a_{A}^{i},a_{A}^{j}]=[b_{A}^{i},b_{A}^{j}]=0, i,j,\displaystyle\quad\forall i,j,
    [aAi,cAs]=[bAi,cAs]=0,\displaystyle\,[a_{A}^{i},c_{A}^{s}]=[b_{A}^{i},c_{A}^{s}]=0, s,i,\displaystyle\quad\forall s,i,
    [cAs,cAt]=0,\displaystyle\,[c_{A}^{s},c_{A}^{t}]=0, s,t.\displaystyle\quad\forall s,t.

    Equivalently, in the standard basis (50), we have a block-diagonal JJ:

    J=(0110)ma(0)mc.J=\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right)^{\oplus m_{a}}\oplus\left(\begin{array}[]{c}0\end{array}\right)^{\oplus m_{c}}. (54)
Proof.

Note that the matrix JJ, defined in (38) can also be treated as a skew-symmetric matrix over F2F_{2}. (On F2F_{2}, we have 1+1=01+1=0.) We can apply the standard method to bring it to the standard form (54). The procedure is to applying a sequence of pairs of row and column operations, where the column operation is similar to the row operation. This sequence of operations is suitable for our purpose because it corresponds to a sequence of redefinition of stabilizer generators. This completes the proof. ∎

Lemma A.5.

In terms of the number mam_{a} defined above,

ENA|B(ρAB)=ma.E_{N}^{A|B}(\rho_{AB})=m_{a}. (55)
Proof.

The proof is based on a few simple observations:

  1. 1.

    Let {hABi}i=1mAB\{h_{AB}^{i}\}_{i=1}^{m_{AB}} be the set of generators of stabilizer group 𝒢AB\mathcal{G}_{AB}. Let

    h~ABi(hABi)TA.\displaystyle\tilde{h}_{AB}^{i}\equiv(h_{AB}^{i})^{T_{A}}. (56)

    Then {h~ABi}i=1mAB\{\tilde{h}_{AB}^{i}\}_{i=1}^{m_{AB}} generates another stabilizer group, which we may denote as 𝒢~AB\tilde{\mathcal{G}}_{AB}.

  2. 2.

    A simple property of partial transpose:

    [(λAλB)(μAμB)]TA=(μATAλATA)(λBμB).[(\lambda_{A}\otimes\lambda_{B})\cdot(\mu_{A}\otimes\mu_{B})]^{T_{A}}=(\mu_{A}^{T_{A}}\cdot\lambda_{A}^{T_{A}})\otimes(\lambda_{B}\cdot\mu_{B}). (57)
  3. 3.

    We must have

    (hABihABj)TA=±h~ABih~ABj(h^{i}_{AB}h^{j}_{AB})^{T_{A}}=\pm\tilde{h}^{i}_{AB}\tilde{h}^{j}_{AB} (58)

    because that each stabilizer generator is a tensor product of factors acting on AA and BB. We obtain “++” if hAih^{i}_{A} and hAjh^{j}_{A} commute and we obtains “-” if hAih^{i}_{A} and hAjh^{j}_{A} anti-commute.

With these simple observations, we find that the density matrix ρAB\rho_{AB}, written the standard basis (50) as

ρAB=12LABi=1ma(1+ai+bi+aibi)s=1mc(1+cs),\rho_{AB}=\frac{1}{2^{L_{AB}}}\prod_{i=1}^{m_{a}}(1+a^{i}+b^{i}+a^{i}b^{i})\prod_{s=1}^{m_{c}}(1+c^{s}), (59)

is of the following form after the partial transpose:

ρABTA=12LABi=1ma(1+a~i+b~ia~ib~i)s=1mc(1+c~s).\rho_{AB}^{T_{A}}=\frac{1}{2^{L_{AB}}}\prod_{i=1}^{m_{a}}(1+\tilde{a}^{i}+\tilde{b}^{i}-\tilde{a}^{i}\tilde{b}^{i})\prod_{s=1}^{m_{c}}(1+\tilde{c}^{s}). (60)

Because different stabilizer generators {a~i,b~j,c~s}\{\tilde{a}^{i},\tilde{b}^{j},\tilde{c}^{s}\} can independently take ±1\pm 1, one can easily verify (55). This completes the proof. ∎

Appendix B Data collapse details

Near the critical point (ppc)(p\approx p_{c}), we expect the physical quantities of interest, e.g., entanglement entropy, mutual information and entanglement negativity to scale as a function of (ppc)L1/ν(p-p_{c})L^{1/\nu}. Here the dimensionless number ν\nu as related to the correlation length by

ξ|ppc|ν.\xi\sim|p-p_{c}|^{-\nu}. (61)

The function can be different for different quantities.

To calculate pcp_{c} and ν\nu, we perform data collapse for the mutual information IA,BI_{A,B} and the entanglement negativity ENA|BE_{N}^{A|B} in the random Clifford circuit model with measurements:

IA,B=f((ppc)L1/ν)\displaystyle I_{A,B}=f((p-p_{c})L^{1/\nu}) (62)
ENA|B=f~((ppc)L1/ν),\displaystyle E_{N}^{A|B}=\widetilde{f}((p-p_{c})L^{1/\nu}), (63)

where we have fixed the two intervals to be antipodal regions with length |A|=|B|=L/8|A|=|B|=L/8 whereas f()f(\cdot) and f~()\widetilde{f}(\cdot) are functions.

We further perform data collapse for random Haar circuit model with measurements. This time, we consider the difference of half-chain entanglement entropy:

S1(p)S1(pc)=g((ppc)L1/ν),S_{1}(p)-S_{1}(p_{c})=g((p-p_{c})L^{1/\nu}), (64)

where g()g(\cdot) is a function.

By following the protocol in Ref. [51], we find that pc=0.16(0.26),ν=1.07(1.35)p_{c}=0.16(0.26),\nu=1.07(1.35) in the Clifford (Haar) case. The results for pcp_{c} in both models are consistent with Ref. [23] and [21]. For the critical exponent ν\nu, our result in the Clifford circuit model is close to that of Ref. [23], but our results of Haar random circuit model differs with ν=2.01\nu=2.01 in Ref. [21]. This is understandable, as in the random Haar case the finite size effect is more significant.

Refer to caption
Refer to caption
Refer to caption
Figure 6: (Top and middle) Data collapse for the mutual information IA,BI_{A,B} and logarithmic negativity ENA|BE_{N}^{A|B} at the critical point pc=0.16p_{c}=0.16 for the hybrid Clifford circuit model. Here we fix the two intervals to be antipodal regions with length |A|=|B|=L/8|A|=|B|=L/8. Thus, the cross-ratio is fixed at η=sin2(π8)0.146\eta=\sin^{2}(\frac{\pi}{8})\approx 0.146. (Bottom) Data collapse for the half-chain entanglement entropy S1(p)S1(pc)S_{1}(p)-S_{1}(p_{c}) for the random Haar model.

References

  • Kitaev and Preskill [2006] A. Kitaev and J. Preskill, Phys. Rev. Lett.  96, 110404 (2006), eprint hep-th/0510092.
  • Levin and Wen [2006] M. Levin and X.-G. Wen, Phys. Rev. Lett.  96, 110405 (2006), eprint cond-mat/0510613.
  • Ryu and Takayanagi [2006] S. Ryu and T. Takayanagi, Phys. Rev. Lett.  96, 181602 (2006), eprint hep-th/0603001.
  • Amico et al. [2008] L. Amico, R. Fazio, A. Osterloh, and V. Vedral, Reviews of Modern Physics 80, 517 (2008), eprint quant-ph/0703044.
  • Zeng et al. [2019] B. Zeng, X. Chen, D.-L. Zhou, and X.-G. Wen, Quantum information meets quantum matter (Springer, 2019).
  • Eisert [2006] J. Eisert, Ph.D. thesis, - (2006).
  • Horodecki et al. [2009] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Reviews of Modern Physics 81, 865 (2009), eprint quant-ph/0702225.
  • Das et al. [2017] S. Das, T. Chanda, M. Lewenstein, A. Sanpera, A. S. De, and U. Sen, arXiv e-prints arXiv:1701.02187 (2017), eprint 1701.02187.
  • Groisman et al. [2005] B. Groisman, S. Popescu, and A. Winter, Phys. Rev. A 72, 032317 (2005), eprint quant-ph/0410091.
  • Wolf et al. [2008] M. M. Wolf, F. Verstraete, M. B. Hastings, and J. I. Cirac, Phys. Rev. Lett.  100, 070502 (2008), eprint 0704.3906.
  • Vidal and Werner [2002] G. Vidal and R. F. Werner, Phys. Rev. A 65, 032314 (2002), eprint quant-ph/0102117.
  • Calabrese et al. [2012] P. Calabrese, J. Cardy, and E. Tonni, Phys. Rev. Lett.  109, 130502 (2012), eprint 1206.3092.
  • Calabrese et al. [2013] P. Calabrese, J. Cardy, and E. Tonni, Journal of Statistical Mechanics: Theory and Experiment 2013, 02008 (2013), eprint 1210.5359.
  • Marvian [2017] I. Marvian, Physical Review B 95, 045111 (2017), eprint 1307.6617.
  • Lee and Vidal [2013] Y. A. Lee and G. Vidal, Phys. Rev. A 88, 042318 (2013), eprint 1306.5711.
  • Wen et al. [2016] X. Wen, G. Y. Cho, P. L. S. Lopes, Y. Gu, X.-L. Qi, and S. Ryu, Phys. Rev. B 94, 075124 (2016), eprint 1605.07199.
  • Shapourian et al. [2017] H. Shapourian, K. Shiozaki, and S. Ryu, Phys. Rev. B 95, 165101 (2017), eprint 1611.07536.
  • Lu et al. [2019] T.-C. Lu, T. H. Hsieh, and T. Grover, arXiv e-prints arXiv:1912.04293 (2019), eprint 1912.04293.
  • Zou et al. [2021] Y. Zou, K. Siva, T. Soejima, R. S. K. Mong, and M. P. Zaletel, Phys. Rev. Lett.  126, 120501 (2021), eprint 2011.11864.
  • Li et al. [2018] Y. Li, X. Chen, and M. P. A. Fisher, Phys. Rev. B 98, 205136 (2018), eprint 1808.06134.
  • Skinner et al. [2019] B. Skinner, J. Ruhman, and A. Nahum, Physical Review X 9, 031009 (2019), eprint 1808.05953.
  • Chan et al. [2019] A. Chan, R. M. Nandkishore, M. Pretko, and G. Smith, Phys. Rev. B 99, 224307 (2019), eprint 1808.05949.
  • Li et al. [2019] Y. Li, X. Chen, and M. P. A. Fisher, Phys. Rev. B 100, 134306 (2019), eprint 1901.08092.
  • Choi et al. [2020] S. Choi, Y. Bao, X.-L. Qi, and E. Altman, Phys. Rev. Lett.  125, 030505 (2020), eprint 1903.05124.
  • Jian et al. [2020] C.-M. Jian, Y.-Z. You, R. Vasseur, and A. W. W. Ludwig, Phys. Rev. B 101, 104302 (2020), eprint 1908.08051.
  • Bao et al. [2020] Y. Bao, S. Choi, and E. Altman, Phys. Rev. B 101, 104301 (2020), eprint 1908.04305.
  • Szyniszewski et al. [2019] M. Szyniszewski, A. Romito, and H. Schomerus, Phys. Rev. B 100, 064204 (2019), eprint 1903.05452.
  • Gullans and Huse [2019a] M. J. Gullans and D. A. Huse, arXiv e-prints arXiv:1905.05195 (2019a), eprint 1905.05195.
  • Gullans and Huse [2019b] M. J. Gullans and D. A. Huse, arXiv e-prints arXiv:1910.00020 (2019b), eprint 1910.00020.
  • 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), eprint 1911.00008.
  • 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), eprint 2001.11428.
  • Fan et al. [2020] R. Fan, S. Vijay, A. Vishwanath, and Y.-Z. You, arXiv e-prints arXiv:2002.12385 (2020), eprint 2002.12385.
  • Li et al. [2020] Y. Li, X. Chen, A. W. W. Ludwig, and M. P. A. Fisher, arXiv e-prints arXiv:2003.12721 (2020), eprint 2003.12721.
  • Nahum et al. [2017] A. Nahum, J. Ruhman, S. Vijay, and J. Haah, Physical Review X 7, 031016 (2017), eprint 1608.06950.
  • Życzkowski et al. [1998] K. Życzkowski, P. Horodecki, A. Sanpera, and M. Lewenstein, Phys. Rev. A 58, 883 (1998), eprint quant-ph/9804024.
  • Calabrese et al. [2009] P. Calabrese, J. Cardy, and E. Tonni, Journal of Statistical Mechanics: Theory and Experiment 2009, 11001 (2009), eprint 0905.2069.
  • Calabrese et al. [2011] P. Calabrese, J. Cardy, and E. Tonni, Journal of Statistical Mechanics: Theory and Experiment 2011, 01021 (2011), eprint 1011.5482.
  • Headrick [2010] M. Headrick, Phys. Rev. D 82, 126010 (2010), eprint 1006.0047.
  • De Nobili et al. [2015] C. De Nobili, A. Coser, and E. Tonni, Journal of Statistical Mechanics: Theory and Experiment 2015, 06021 (2015), eprint 1501.04311.
  • Gottesman [1998] D. Gottesman, arXiv e-prints quant-ph/9807006 (1998), eprint quant-ph/9807006.
  • Christandl and Winter [2004] M. Christandl and A. Winter, Journal of Mathematical Physics 45, 829 (2004), eprint quant-ph/0308088.
  • Gurarie and Ludwig [2005] V. Gurarie and A. W. W. Ludwig, Conformal Field Theory at Central Charge c = 0 and Two-Dimensional Critical Systems with Quenched Disorder (2005), pp. 1384–1440.
  • Pearce et al. [2006] P. A. Pearce, J. Rasmussen, and J.-B. Zuber, Journal of Statistical Mechanics: theory and experiment 2006, P11017 (2006).
  • Dupic et al. [2018] T. Dupic, B. Estienne, and Y. Ikhlef, SciPost Physics 4, 031 (2018), eprint 1709.09270.
  • Jacobsen and Saleur [2008] J. L. Jacobsen and H. Saleur, Phys. Rev. Lett.  100, 087205 (2008), eprint 0711.3391.
  • Vasseur et al. [2019] R. Vasseur, A. C. Potter, Y.-Z. You, and A. W. W. Ludwig, Phys. Rev. B 100, 134203 (2019), eprint 1807.07082.
  • Nahum et al. [2018] A. Nahum, S. Vijay, and J. Haah, Physical Review X 8, 021014 (2018), eprint 1705.08975.
  • Sang et al. [2020] S. Sang, Y. Li, T. Zhou, X. Chen, T. H. Hsieh, and M. P. A. Fisher, arXiv e-prints arXiv:2012.00031 (2020), eprint 2012.00031.
  • Bravyi et al. [2006] S. Bravyi, D. Fattal, and D. Gottesman, Journal of Mathematical Physics 47, 062106 (2006), eprint quant-ph/0504208.
  • Aaronson and Gottesman [2004] S. Aaronson and D. Gottesman, Phys. Rev. A 70, 052328 (2004), eprint quant-ph/0406196.
  • Houdayer and Hartmann [2004] J. Houdayer and A. K. Hartmann, Phys. Rev. B 70, 014418 (2004).