You are on page 1of 27

Measurement-induced phase transitions on dynamical quantum trees

Xiaozhou Feng,1 Brian Skinner,1 and Adam Nahum2


1
Department of Physics, The Ohio State University, Columbus, Ohio 43202, USA
2
Laboratoire de Physique de l’École Normale Supérieure,
ENS, Université PSL, CNRS, Sorbonne Université,
Université Paris-Diderot, Sorbonne Paris Cité, Paris, France.
(Dated: October 17, 2022)
Monitored many-body systems fall broadly into two dynamical phases, “entangling” or “disentan-
gling”, separated by a transition as a function of the rate at which measurements are made on the
arXiv:2210.07264v1 [cond-mat.stat-mech] 13 Oct 2022

system. Producing an analytical theory of this measurement-induced transition is an outstanding


challenge. Recent work made progress in the context of tree tensor networks, which can be related
to all-to-all quantum circuit dynamics with forced (postselected) measurement outcomes. So far,
however, there are no exact solutions for dynamics of spin-1/2 degrees of freedom (qubits) with
“real” measurements, whose outcome probabilities are sampled according to the Born rule. Here
we define dynamical processes for qubits, with real measurements, that have a tree-like spacetime
interaction graph, either collapsing or expanding the system as a function of time. The former
case yields an exactly solvable measurement transition. We explore these processes analytically and
numerically, exploiting the recursive structure of the tree. We compare the case of “real” measure-
ments with the case of “forced” measurements. Both cases show a transition at a nontrivial value of
the measurement strength, with the real measurement case exhibiting a smaller entangling phase.
Both exhibit exponential scaling of the entanglement near the transition, but they differ in the value
of a critical exponent. An intriguing difference between the two cases is that the real measurement
case lies at the boundary between two distinct types of critical scaling. On the basis of our results
we propose a protocol for realizing a measurement phase transition experimentally via an expansion
process.

I. INTRODUCTION exact analytical treatment, except in the limit of infinite


local Hilbert space dimension, where the phase transi-
tion reduces to a problem of the classical geometry of
If the unitary time evolution of a many-body quantum
the circuit. There is an active effort to describe the
system is punctuated by measurements, made at a finite
MPT and other models with related mathematical struc-
rate per local degree of freedom, the resulting dynamics
ture using effective statistical mechanical lattice mod-
can fall into one of two broad classes known as the “en-
els [8, 9, 22, 27–32] or effective “Landau-Ginsburg-like”
tangling” and “disentangling” dynamical phases, or the
field theories [22, 27]. But so far there is no theory that
“weak monitoring” and “strong monitoring” phases. Sep-
can reproduce the numerically observed critical measure-
arating these two phases is a measurement-induced phase
ment rate nor the critical exponents of the MPT in a
transition (MPT), which can be crossed by increasing the
low-dimensional system of qubits.
rate at which measurements are performed [1–23].
This challenge has motivated simplifying the spatial
In the original identification of the MPT [1, 2], the structure of the problem. Recent work examined the
transition was associated with the scaling properties of MPT in all-to-all coupled circuits [7, 22, 23]. In these
the entanglement entropy of a pure state (volume-law models unitary gates are applied (at random, with equal
versus area-law). Subsequent work made clear that there rates) between any pair of spins/qubits, and these uni-
are alternative useful formulations of the MPT. In partic- taries are interspersed with a finite rate of single-spin pro-
ular, one can associate the MPT with a sharp change in jective measurements. Such models take the spin system
the rate at which an initially mixed state is transformed to a mean-field-like limit in which all pairs of spins can
into a pure state by the monitored dynamics [7]. In the interact. The hope is that this limit will enable an ana-
disentangling phase the characteristic timescale for pu- lytical approach that is not available in finite-dimensional
rification is only order-one, while in the entangling phase systems.
purification takes a time that is exponential in the sys- Recent progress in this direction has made use of the
tem size. This purification timescale is also the timescale fact that the quantum circuit for a large all-to-all-coupled
up to which quantum information from the initial state is system has a locally tree-like structure [22], leading to
retained in the final state [6, 20, 22, 24], and is therefore the conjecture that the MPT in the all-to-all circuit co-
related to questions like “do I need to know the initial incides with an entanglement transition in an associated
state in order to deduce the final state, or can I deduce ensemble of tree tensor networks. This correspondence
it from the measurement outcomes alone?”. enables exact results in the case where the measurement
While numerical studies have characterized the MPT outcomes are “forced”, as described below.
in a wide variety of settings (see Refs. [25, 26] for re- In order to characterize the entanglement transition on
views), so far its critical properties have largely evaded an a tree tensor network, one can define an “order param-
2

eter” Zk (defined more precisely below) which quanti-


fies the degree of entanglement between the root and the
leaves of a tree with k generations. In the weak monitor-
ing phase, this order parameter remains nonzero when
the tree’s size diverges. Importantly, the tree structure
allows one to solve for the behavior of Zk as k → ∞ using
a recursion relation, giving analytic results for the critical
measurement rate and the critical vanishing of Zk near
the transition [22]. These results generalize to entangle-
ment transitions in a large class of tree tensor networks.
These developments, as well as the results in [33, 34], FIG. 1. Schematic illustrations of the collapse process (Left)
suggest that tree tensor networks [33, 35–43] are a useful and expansion processes (Right). Each node of the tree rep-
tool for developing a mean-field theory for measurement resents an interaction between two qubits, preceded/followed
and entanglement transitions. by weak measurements. Viewing the tree as a tensor net-
Despite this progress, we do not yet have an exact so- work, the symbol t denotes a local three-legged tensor and T
denotes the full network. (These tensors are defined in the
lution for an MPT for qubits with true measurements,
text.)
even in the context of quantum trees. In Ref. 22, a re-
duction to a tractable tree tensor network was obtained
only for a circuit where measurement outcomes were pre-
half of the qubits at each timestep, we are left with a
determined, instead of being sampled with Born’s rule.
single qubit at time t = k. Interactions and weak mea-
Physically, the former situation corresponds to a proto-
surements also take place during each timestep. In the
col where one runs the dynamics many times and then
“expansion process” (Fig. 1, right) we start with a sin-
discards all realizations except those for which the mea-
gle qubit and (by recruiting further spins) attempt to
surements produce a desired sequence of outcomes (e.g.,
scramble its state into a many-body state of 2t qubits.
all spins are measured in the ↑ state). This protocol
In both cases, we ask whether a maximally mixed initial
was dubbed the “forced measurement” case, and the
state is purified by the monitored dynamics (the strong
resulting transition is called the “forced-measurement-
monitoring phase) or not (the weak monitoring phase).
induced phase transition” (FMPT). A key question, then,
The transition is induced by varying the strength of the
is whether the approach of Ref. 22 can be extended
weak measurements that occur in every timestep. We
to describe real measurements, for which the measure-
focus mainly on the collapse process, where we give ex-
ment outcomes are sampled with the nontrivial, state-
act results for both MPT and FMPT. (Our results imply
dependent probability given by Born’s rule. More gen-
exact results for the expansion process FMPT, but not
erally, one can ask how the MPT and FMPT differ in
the expansion process MPT, as we discuss in more detail
various settings. Do they have different critical points
below.)
and/or different critical exponents? It would be valuable
Let us contrast the “dynamical” trees studied here
to have a model in which exact results can be obtained
with the trees studied in Ref. [22]. In the present work,
for a version of the MPT.
the “generation” coordinate of the tree is simply propor-
In this paper we define a different kind of many-body tional to the physical time coordinate t. By contrast, in
dynamics, the “collapse process”, which has a solvable Ref. [22] trees arose more indirectly, starting from circuit
phase transition for real measurements as well as for dynamics with a fixed number of qubits: in that case the
forced measurements. This solvability allows us to ex- “generation” coordinate of the tree was not equivalent to
amine the difference between real and forced measure- physical time.1 Ultimately, in both cases we must deal
ments. We also define a closely-related “expansion pro- with a random ensemble of tree tensor networks. How-
cess”. Cartoons of these two processes are shown in ever, the fact that we consider real measurements here
Fig. 1. A potentially useful feature of dynamical trees means the probability distribution for this ensemble of
is that they exhibit measurement phase transitions that trees has nontrivial correlations between different nodes.
are experimentally accessible without the need for post- Crucially, the collapse model we introduce maintains
selection [12, 44–48] on measurement outcomes, because a unitary invariance property for the distribution of the
the efficient contractability of tree tensor networks can output states, even for the case of real measurements: as
be exploited to classically process experimental measure- a result of the Haar-randomness of the interaction uni-
ment outcomes. taries, there is no preferred local basis. This invariance
The dynamics we discuss are not conventional quan-
tum circuit dynamics with a fixed number of qubits. In-
stead we define dynamical protocols whose spacetime in-
teraction graph resembles a tree, with the number of ac- 1 In the previous case the generation coordinate represented the
tive qubits either decreasing or increasing in time. In graph distance from an arbitrarily chosen bond of the circuit
the “quantum collapse process” (Fig. 1, left) we begin (i.e. an arbitrary point on the worldline of some spin): see Fig. 4
with 2k  1 qubits, and by measuring and discarding of Ref. 22.
3

allows us to locate the transition analytically by studying ln 2


a recursive map [22]. The invariance also constrains the MPT
universality class of the phase transition [22, 34]. FMPT
We find that the MPT and FMPT have qualitatively
similar critical behavior, with an exponential vanishing
of the order parameter as the transition is approached 0.4

SvN
from the weak-monitoring side. However they have dif-
ferent values of the critical measurement strength, with
the MPT exhibiting a smaller entangling phase than the
FMPT. This difference can be seen in Fig. 2, which shows
the von Neumann entropy SvN of the final state (i.e. the
final, unmeasured spin at the apex of the tree in Fig. 1) 0.0 π π
for a collapse process in the limit of long time (large 4 θcM 3π
8
θcF 2
initial system size). In the strong monitoring phase, θ
SvN → 0 in the limit of long time, while in the weak
monitoring phase SvN remains finite. For both types of
dynamics, SvN is plotted against a parameter θ (defined FIG. 2. The purity of the state of the final spin in a collapse
below) that characterizes the strength of weak measure- process, as characterized by the von Neumann entropy SvN
of the corresponding density matrix. Here we plot the value
ments within each node of the tree; θ → π/4 corresponds
of SvN in the limit of long time (large initial system size) for
to the limit of no measurement while θ → π/2 is the limit both the real measurement (MPT) and forced measurement
of strong, projective measurements. (FMPT) cases. The horizontal axis shows the parameter θ
Interestingly, the MPT and FMPT reside in distinct that defines the strength of the weak measurement within
universality classes, with the MPT (but not the FMPT) each node of the tree (see Sec. II A); larger θ corresponds to
located at the boundary separating two different regimes stronger measurement. The calculated critical values for the
for the phase transition. These two regimes can be MPT and FMPT cases are shown as θcM and θcF , respectively.
thought of as “strong” versus “weak” randomness. This Symbols correspond to results from numerical calculations,
result suggests that there may be slightly different proto- and error bars are smaller than the symbol size.
cols for which the MPT and FMPT exhibit more radical
differences in their critical scaling, for example power-law since the final spin is a 2-state system). This entropy is
versus exponential scaling of the order parameter. a measure of the amount of surviving information, and
The structure of this paper is as follows. In Sec. II we is also an order parameter for the phase transition.
define the dynamical tree models. In Sec. III we study The expansion process is, loosely speaking, the inverse
the forced measurement case using both simulations and process. Initially the “system” consists of a single spin.
theoretical analysis. In Section IV we turn to the case Further spins (initialized in a definite state, |↑i) are re-
of real measurements, and contrast it with the forced cruited at each timestep so that at time t the system
measurement case. In Sec. V we discuss the expansion consists of 2t spins. Again we ask whether a maximally
process, and suggest a protocol for realizing the mea- mixed initial state is purified or not.
surement phase transition on a tree experimentally. We An equivalent way to formulate this question about
conclude in Sec. VI with a summary and a discussion of purification [12] in the expansion process is to start with
potential future work. a state in which the initial spin is maximally entangled
with a “reference” spin. In this case entropy of interest is
also equal to the entanglement, at the final time, between
II. DYNAMICAL TREE MODELS the reference spin and the collection of 2t system spins
(which is again no larger than ln 2).
The quantum dynamics we consider are of two kinds. The detailed dynamics are specified below and include
In the collapse process we start with a large number of unitary interactions and weak measurements. Equiva-
spins (qubits). In each timestep the number of “active” lent models could also be formulated using only projec-
spins is halved, because, after the spins interact in pairs, tive measurements, at the cost of a more complex node
half of them are projectively measured and discarded interaction involving ancilla spins (a trivial way to see
(Fig. 1). This process gives a tree structure in space- this equivalence is by using the physical interpretation of
time, with only a single spin remaining at the top of the a weak measurement in terms of an ancilla2 ).
tree, i.e. at the final time. The question we ask is whether From a formal point of view, the basic objects describ-
the state of this final spin preserves quantum information ing the evolution are tree tensor networks, built up (as
from the initial state [7]. This question may be formal-
ized by initializing the spins in the lowest layer of the tree
in the maximally mixed state and asking whether the fi- 2 A weak measurement can be achieved by allowing the spin of
nal spin is in a pure state, or whether it is in a mixed interest to interact for a finite time with an ancilla degree of
state with a nonzero entropy (which can be at most ln 2, freedom, and then projectively measuring the ancilla [49].
4

The Haar-randomness of the two-site unitaries means


that it does not matter whether we choose to make all
the measurements in the Sz basis, or whether we choose
a new random basis for each measurement. These two
measurement protocols are related by a “gauge transfor-
mation” which involves rotating the measurement bases
and making compensating rotations of the 2-site unitaries
(see Appendix A for more details).3 For notational con-
venience later on, we will take each weak measurement
to be in an independent random basis, but the strong
measurements to be in the Sz basis.
The unmeasured spin that remains after (i)–(iii) is then
used as an input for another node of the tree, unless it is
the final spin at the apex of the tree. The resulting tree
is depicted in Fig. 3. The strength of the weak measure-
ment in step (i) is the major parameter of the model.
FIG. 3. An illustration of the tree model studied in this paper These ingredients may be formalised as follows. Each
(for the collapse process), drawn here with k = 3 generations. weak measurement is described using a Kraus operator
Each line represents the world line of one of the 2k spins in
the system. The red K blocks (defined by Eq. 1) represent Kσ = cos θ1 + (sin θ − cos θ) u |σi hσ| u† , (1)
single-spin weak measurements. The blue U blocks represent
independent Haar-random, two-spin unitary operators. The where σ =↑ or ↓ labels the measurement outcome, |σi de-
red bras are associated with projective (strong) single-spin notes an Sz eigenstate, and u is a Haar-random unitary
measurements. A single node of the tree (a circle in Fig. 1), rotation that randomizes the measurement basis. (An
comprising all three operations, is indicated by the orange independent random basis is chosen for each weak mea-
dashed box.
surement, but to avoid clutter in our notation we will
leave the dependence of Kσ on u implicit.) If a given
shown in Fig. 3 for the collapse process) from: unitary spin has the initial 2 × 2 density matrix ρ, then after
gates U ; Krauss operators K (representing weak mea- a weak measurement with outcome σ the spin has the
surements, see below); and kets/bras associated with the normalised density matrix
spins that are recruited/discarded in each timestep. The
Kσ ρKσ†
set of allowed tree tensor networks is identical for both . (2)
the collapse and expansion processes, just with the time trKσ ρKσ†
coordinate reversed. In both cases the order parameter
is a simple property of the singular value decomposition If the weak measurement is a “true” measurement, the
of the tree (for a decomposition separating the “trunk” probability of the outcome σ is equal to
from the “leaves”). However, the probability distribution
pσ = trKσ ρKσ† . (3)
on this set of tree tensor networks depends crucially on
the physical interpretation of the dynamics, because ap- For a forced measurement, we fix the outcome to σ =↑,
plying Born’s rule (the usual quantum mechanical prob- without loss of generality. See the following subsection
ability for measurement outcomes) leads to correlations for a comment on the physical interpretation of forced
between the constituent elements of the tensor network. measurements.
We now describe the dynamics more carefully.
The strength of the weak measurement is defined by
the parameter θ,
A. The collapse process: basic “node” θ ∈ [π/4, π/2] . (4)

The basic ingredient of the collapse process (i.e. a node When θ = π/4 the Kraus operator is proportional to the
of the tree) is a random operation that takes a state of identity and no measurement occurs, while when θ = π/2
two spins as the input and gives a state of a single spin the Kraus operator is a projection operator, giving a con-
as output. The node consists of the following sequence ventional projective measurement. For both the MPT
of three steps: and the FMPT we will find that the critical value of θ
(i) a weak measurement is applied independently to lies in the interior of this interval, with θ = π/4 being in
each of two spins along the lower bonds,
(ii) a Haar-random, two-spin unitary operator U is ap-
plied to the two spins, and
3
(iii) one of the two spins undergoes a projective mea- Here we assume that the initial state is invariant under single-site
surement. unitary rotations, as for the maximally mixed case.
5

the entangling phase and θ = π/2 in the disentangling B. Collapse process as a tensor network
phase.
Let the initial density matrices of the (uncorrelated) The full collapse process can be described with the ten-
spins involved in the node be ρ1 and ρ2 . Then the action sor network T which is shown in Fig. 3. This tensor net-
of the weak measurements is work can be viewed as a 2×M rectangular matrix, where
(1) (1)† (2) (2)† the number of rows, 2, is the Hilbert space dimension of
Kσ1 ρ1 Kσ1 Kσ2 ρ2 Kσ2 the final remaining spin (the topmost bond of T ) and
ρ1 ⊗ ρ2 −→ ρ0 ≡ (1) (1)†
⊗ (2) (2)†
, (5) k
trKσ1 ρ1 Kσ1 trKσ2 ρ2 Kσ2 the number of columns, M = 22 , is the Hilbert space
k
dimension of the 2 initial spins (the lowermost bonds).
with the appropriate probabilities for the outcomes σ1 , T depends on the various random variables as follows.
σ2 . (We write the superscripts (1) and (2) for the Kraus First, there are the various random unitaries. We de-
(1)
operators because K↑ — for example — differs from note the complete set of all the unitaries appearing in T
(2)
K↑ as a result of the independent choices of random by U. For each node r there is a two-site random unitary
(1)
measurement basis for each spin). Next, in step (ii) we Ur , together with two single-site random unitaries, ur
(2)
generate a 4 × 4 Haar-random unitary matrix U and act and ur , that appear inside the Kraus operators and set
on the combined state of the two spins so that the new the weak measurement bases. Second, T depends on the
4 × 4 density matrix is lists, denoted mW and mS , of measurement outcomes
for the weak and strong measurements respectively. We
ρ00 = U ρ0 U † . (6) will sometimes make the dependence on the measure-
ment outcomes explicit, T = T (mW , mS ), but we will
Finally, in step (iii) we perform a projective measurement leave the dependence of T on the various unitaries im-
of one spin (taken here to be spin 2, without loss of gen- plicit. We will do the same for the local node tensors t
erality) in the S z basis. For measurement outcome σ20 described below.
(which can again either be selected with Born’s rule, or The tensor network T is built by contracting together
postselected to enforce σ20 =↑) the corresponding projec- three-index tensors (tr )abc , where r labels a trivalent node
tion operator is Pσ20 = 1 ⊗ (|σ20 i hσ20 |), and the final 2 × 2 of the tensor network (from now on we will suppress this
density matrix for the first spin after all three steps is subscript) and a, b, c are the bond indices. The node
tensor is constructed as follows. First, we define
tr2 Pσ20 ρ00 Pσ20
ρ000 = . (7) t(σ1 , σ2 ) = U (Kσ(1)
e
1
⊗ Kσ(2)
2
), (11)
trPσ20 ρ00 Pσ20
which is an operator acting on two spins. We write its
Here tr2 denotes the partial trace over the second spin, components as e tab
cd , where the positions of the indices re-
while tr denotes the trace over both spins. flects their geometrical position in the tensor network:
For real measurements, the joint probability for the indices (a, b) are for the upper bonds of the tensor,
p(σ1 , σ2 , σ20 ), given the initial state and the random uni- and (c, d) are for the lower bonds. Viewed as a matrix,
taries, can be written as the product of the probabilities the multi-index (a, b) makes up the row index of e t and
for the earlier measurements and the conditional proba- (c, d) makes up the column index.
bility for the later one: The strong measurement outcome fixes the value of the
0 index b, so the desired three-index tensor, conditioned on
p(σ1 , σ2 , σ20 ) = p(2 ) (σ20 |σ1 , σ2 ) p(1) (σ1 ) p(2) (σ2 ). (8) all three measurement outcomes, is:
0
Using Eq. 3, and p(2 ) (σ20 |σ1 , σ2 ) = trPσ20 ρ00 Pσ20 , gives [t(σ1 , σ2 , σ20 )]acd = e
a,σ 0
t(σ1 , σ2 )cd 2 . (12)
p(σ1 , σ2 ,σ20 ) =
h   i If we take the 2k spins that comprise the initial state
tr Pσ20 U Kσ(1) ρ K (1)†
1 σ1 ⊗ K (2)
ρ
σ2 2 σ2K (2)†
U †
P σ2 .
0 to be in a maximally mixed state, so that the initial den-
1
sity matrix of the system is proportional to the identity
Anticipating the notation of the next section, the con- matrix, ρ = 1/M , then the final state of the last spin
ditional probability can be written more conveniently in can be written as (again we make the dependence on the
terms of a node tensor t, measurement outcomes explicit):

p(σ1 , σ2 , σ20 ) = tr t(σ1 , σ2 , σ20 )(ρ1 ⊗ ρ2 )t(σ1 , σ2 , σ20 )† . (9) T (mW , mS )T (mW , mS )†
ρf (mW , mS ) = . (13)
tr T (mW , mS )T (mW , mS )†
Similarly, the output density matrix is
The matrix multiplication in T T † contracts all the bot-
t(σ1 , σ2 , σ20 )(ρ1 ⊗ ρ2 )t(σ1 , σ2 , σ20 )† tom bonds of the tensor network T (Fig. 3) with the
ρ000 = . (10)
tr t(σ1 , σ2 , σ20 )(ρ1 ⊗ ρ2 )t(σ1 , σ2 , σ20 )† corresponding bonds of T † , leaving a 2 × 2 matrix.
6

We now consider the probability distribution for the C. Characterizing the transition
tensor network T for protocols involving either forced or
true measurements. We adopt the perspective mentioned in the introduc-
In all cases, the unitaries in the set U are drawn in- tion, and introduced in Ref. [7], of whether a maximally-
dependently from the Haar distribution. In the forced mixed initial state is purified by the dynamics. This is
measurement case, this set of unitaries is sufficient to fix equivalent to studying the singular value decomposition
T entirely. Using the unitaries we have drawn, we can of the tree tensor network/rectangular matrix T .
form the local node tensors t = t(↑, ↑, ↑).
As above, the 2k spins in the collapse process are ini-
Note that in this forced measurement case the local t
tially in a maximally mixed state. We ask whether the
tensors for the nodes are independently random, i.e. they
density matrix ρf of the unmeasured spin at the apex
are uncorrelated between nodes.4 In the true measure-
of the tree is in a pure state or a mixed state. In its
ment case the local measurement outcomes, and therefore
eigenbasis, ρf may be written
the local tensors t = t(σ1 , σ2 , σ20 ), are correlated.
For true measurements, the tree T (mW , mS ) is deter-  
mined by both U and the measurement trajectory. The 1−Z 0
ρf = , 0 ≤ Z ≤ 1/2. (16)
probability distribution for T (mW , mS ) is the product 0 Z
of the Haar distribution for U and the conditional prob-
ability (given U) of obtaining the trajectory (mW , mS ). The value of Z, the smaller of the two eigenvalues of
The probability of such a trajectory may be written as ρf , determines how mixed the state is. Z = 0 cor-
responds to a pure state, while Z = 1/2 describes
1 a maximally mixed state. In the near-critical regime,
p(mW , mS ) = tr T (mW , mS )T (mW , mS )† . (14)
M where Z is small, the Rényi entropies are approximately
In practise, rather than using this expression for the full SvN ≡ S1 =' Z ln Z −1 , and Sn ' n(n−1)−1 Z for n > 1.
trajectory, it will be more convenient to think of the pro- We use the notation Zk to denote the smaller eigen-
cess as a Markov process, in which at each time step the value of a tree of k generations. In the limit of large k,
density matrices are acted on in the manner described in Zk yields an order parameter for the MPT or FMPT:
the previous section. the entangling phase is defined by the fact that the typ-
We have distinguished the case where all measure- ical (or the average) value of Zk is nonzero in the limit
ments are true from the case where all measuremements k → ∞. As discussed above, Z is closely related to the
are forced. We will also mention a mixed case, where entropies, so that having finite Z in the limit k → ∞ is
the weak measurements are true measurements, but the equivalent to having a nonzero entropy in the limit of a
strong measurements are postselected to be ↑. In this large tree. We have already shown numerical data for the
case, the probability for a given tree is determined by average entropy hSvN i in Fig. 2, illustrating the critical
conditioning the probability Eq. 14 on having mS = 0, vanishing at the relevant critical value of θ.
where 0 is the list (↑, ↑, . . . , ↑) of length 2k − 1. This The evolution of the distribution of Zk as a function of
process gives5 k can be described recursively, as detailed in the following
1 sections. For an analytical treatment it will be convenient
pmixed (mW ) = tr T (mW , 0)T (mW , 0)† (15) to focus on the typical value of Zk [50], defined by6
2
(where the nontrivial probability distribution is only for
mW , since mS is fixed). As we will discuss, the distri- ln Zktyp = hln Zk i, (17)
bution of weak measurement outcomes on the right-hand
side of Eq. 15 is also precisely the one relevant to the instead of the mean value hZk i. This distinction is im-
expansion process with true (weak) measurements. portant because Zk may have a broad statistical distribu-
tion for trees of size k  1, so that hZk i is dominated by
rare samples and is much larger than the typical value.
4 Physically, we could sample the forced measurement trees using (Nevertheless, limk→∞ Zktyp and limk→∞ hZk i are both
the following procedure. We first choose U, then attempt to run nonzero precisely in the entangling phase.)
the dynamics. If the dynamics produces a “wrong” measurement To recap, the phenomenology of the transition is that
outcome, i.e. ↓, we discard that run and re-run the dynamics (in
fact it is sufficient to re-run the relevant subtree) using the same
in the entangling phase the typical value Zktyp tends to
U. We repeat this process until we get an all-↑ measurement a positive constant in the limit k → ∞, indicating that
record for the given U. Then we resample U and repeat. Note a nonzero entropy is retained in the final state, while in
that this protocol does not bias the distribution of U: all unitaries the disentangling phase, Zktyp → 0 as k → ∞.
are Haar-distributed.
5 In more detail: pmixed (mW ) is the conditional
probability Prob(mW |mS = 0). In terms of
p(mW , mS ) in Eq. 14, Pthis conditional probability is
pmixed (mW ) = p(mW , 0)/( m0 p(m0W , 0)). Using (14) 6
W In Ref. [22], care was needed to take the average in Eq. 17 over
P †
and repeatedly using the relation σ Kσ Kσ = 1 for the Kraus only nonzero values of Zk . In the model we study here, however,
operators, the denominator simplifies to 21−M . the value of Zk is always nonzero for all finite k unless θ = π/2.
7

D. Expansion process A. Linearized Recursion Relation

The expansion process at time t = k (Fig. 1, Right) is In order to identify a recursion relation for the prob-
characterized by a tensor network that relates the state ability distribution of the smaller eigenvalue Z (see
of a single initial spin to the state of all 2k spins at the Eq. 16), we imagine the process of constructing a tree
final time. (This final state includes the 2k − 1 spins that with k + 1 generations from two smaller subtrees, each
were recruited during the k timesteps.) with k generations. We denote the probability distribu-
tion of the smaller eigenvalue for a tree of k generations
The relevant tensor networks are related to the ones
by µk (Z).
in the collapse process, defined in Sec. II B, by reversing
Suppose that the two subtrees have final states that
the direction of the arrow of time.
are described by density matrices ρk,1 and ρk,2 , respec-
For the case of forced measurements, where all mea- tively, with corresponding smaller eigenvalues Zk,1 and
surements are postselected to be ↑, the expansion pro- Zk,2 . Since the two subtrees are statistically indepen-
cess is precisely equivalent to the collapse process except dent, each of these eigenvalues is drawn independently
for this time reversal. A node in the expansion process from the distribution µk . Combining these trees into a
which “recruits” a spin in the state |↑i can be viewed larger tree with k + 1 generations via the “node” opera-
as the time reversal of a node in the collapse process tion described in Sec. II A, with the appropriate proba-
that discards a spin after a forced measurement in the bilities for the unitaries and measurements, gives a new
state |↑i. More precisely, the relevant ensemble of ten- final density matrix ρk+1 whose smaller eigenvalue Zk+1
sor networks is the exactly same between the two forced is distributed according to µk+1 .
measurement processes (with the probability distribution In considering this process of combining subtrees, it
given simply by the Haar distributions for the unitaries). is in fact sufficient to take ρk,1 and ρk,2 to be diago-
Since the FMPT is a property of the ensemble of tensor nal and of the form in Eq. 16. This simplification arises
networks, it follows that the expansion process has all because the unitary transformations required to achieve
the same critical properties as the collapse process in the this diagonal form can be absorbed into the Haar-random
FMPT case. unitaries involved in the node tensor, without changing
In the expansion process with true measurements, the probability distribution of these operations. This is a
the weak measurement outcomes mW are sampled with crucial simplification from Haar randomness: it means we
Born’s rule (there are no strong measurements in the ex- can deal with a recursion relation only for the eigenvalue
pansion process). This process is not equivalent to the Z, without having to keep track of how the associated
collapse process with only true measurements. Instead, eigenvectors of the density matrix evolve.
writing the relevant probability pexpansion (mW ) in terms In a given instance, the new eigenvalue Zk+1 depends
of T shows that pexpansion (mW ) is equal to the probability on the values of Zk,1 and Zk,2 , as well as on the variables
pmixed (mW ) of the “mixed” collapse process in Eq. 15, in the node (U , and the random basis rotations in K (1)
where the weak measurements are true but the strong and K (2) ). For example, if Zk,1 = Zk,2 = 0, then we
ones are forced. necessarily have Zk+1 = 0, since the action of the node
In our discussion of the MPT in Sec. IV we focus on the always takes two independent pure states into a single
collapse process. We comment in Sec. V on an efficient pure state.
experimental protocol for detecting the transition in an In order to understand the critical properties near the
expansion process. transition, it is sufficient to consider the case where Zk is
very small compared to unity (either vanishing at large
k in the disentangling phase, or tending to a very small
constant just inside the entangling phase). Much of the
key information is contained in the leading order (linear)
III. FORCED MEASUREMENTS
expansion of Zk+1 in terms of Zk,1 and Zk,2 :

Before considering real measurements, in this section Zk+1 = A1 Zk,1 + A2 Zk,2 + O(Zk2 ). (18)
we examine the case of forced measurements, as these can For a given node, the coefficients A1 and A2 can be ex-
be understood by a straightforward application of the ap- pressed in terms of the three-index tensor t defined in
proach in Ref. 22. We briefly recapitulate this approach, Sec. II B, which makes up a local node of the tree tensor
then give the resulting critical properties for the FMPT. network. For the true measurement problem, this tensor
As mentioned above, in the forced measurement case all depends on the measurement outcomes at the node,
measurement outcomes are fixed in advance to give the
↑ state in the corresponding measurement basis. t = t(σ1 , σ2 , σ20 ), (19)
Below we discuss the collapse process FMPT. However while for the forced measurement problem these are fixed,
the results below also apply for the expansion process and
FMPT, for the reason described in Sec. II D: the same
ensemble of tensor networks describes both problems. t = t(↑, ↑, ↑). (20)
8

for large k, i.e. in the nature of the probability distribu-


tion µk after many recursive steps.
The linearized recursion for Z (Eq. 18) implies a re-
cursion relation for Gk (x) that can be expressed through
the coefficients A1 and A2 as:
Gk+1 (x) = hGk (x − ln A1 )Gk (x − ln A2 )i, (26)
where the remaining average is over A1 and A2 . Note
that the probability distribution of A1,2 depends on the
measurement strength parameter θ of our model.
It is convenient to view k as the time and x as a kind
of position for a travelling wave [50], as suggested by
Fig. 4. Then Eq. 25 describes a wave front that is located
at x = ln Zktyp and which propagates in “space” with a
FIG. 4. Schematic illustration of the generating function velocity vθ that may be either positive or negative. The
Gk (x) (solid blue curve). The value of ln Zktyp corresponds critical point θc corresponds precisely to the value for
to the position of the wave front. In the disentangling phase, which vθ vanishes [22].
this front moves toward negative x with increasing time (in- In the disentangling phase, the wave front propagates
creased k). to the left (vθ < 0) at large times k, meaning that the
typical value of Zk decays exponentially, Z typ ∼ e−|vθ |k .
In the entangling phase, on the other hand, the linearized
The tensor t is therefore defined by the Haar-random
treatment gives a traveling wave that propagates to the
2-site interaction unitary and the two Haar-random 1-
right (vθ > 0). Since a rightmoving wave would corre-
site unitaries appearing in the Kraus operators (cf.
spond to Z typ growing indefinitely, the nonlinear terms in
Eqs. 11, 12):
Eq. 18 become important in the entangling phase: they
(1) (2) a↑
tacd = U (K↑ ⊗ K↑ ) cd .

(21) halt the front at finite value of ln Zktyp as discussed below
in Sec. III B. However, our first task is to understand the
In terms of t, Eq. 7 becomes linearized problem, which is captured by Eq. 26: from
this linearized recursion we can identify the location of
t(ρk,1 ⊗ ρk,2 )t† the critical point.
ρf = , (22)
tr t(ρk,1 ⊗ ρk,2 )t† The travelling wave ansatz is
where tacd is treated as a 2 × 4 rectangular matrix whose Gk (x) = G(λ) x − vθ (λ)k .

(27)
row index is a and whose column index is (c, d).
From this expression for ρf we can write the expres- Here λ ≥ 0 parameterizes a family of possible travelling
sion for trρ2f , and the latter can be expanded in Zk+1 wave solutions: it will be necessary to choose the cor-
as trρ2f ' 1 − 2Zk+1 . Explicitly evaluating the left-hand rect solution that is appropriate to the initial conditions.
The value of λ is defined through the exponential de-
side gives the random coefficients in (18) as:
cay of G(λ) at large argument (see Ref. 50 and references
t11 t21 − t121 t211 2
1 2 therein):
A1 = 2 , (23)
(|t111 |2 + |t211 |2 ) G(λ) (u) ' 1 − e−λu . (28)
t11 t12 − t112 t211 2
1 2
A2 = Using this asymptotic expression and the recursion rela-
2 . (24)
(|t111 |2 + |t211 |2 ) tion for Gk (x) (Eq. 26) gives the following result for the
wave front velocity:
In order to understand the evolution of Zk with in-
creasing k, we define the generating function [50] 1
ln Aλ1 + Aλ2



vθ (λ) = (29)
−x
λ
Gk (x) = hexp(−e Zk )i. (25)
where again the average is over A1,2 as defined by Eqs. 23
Here, h...i indicates the average over Zk with distribution and 24, or in other words over the random unitary op-
µk . A heuristic way to think of Gk (x) is as a smeared ver- erations that define the node tensor t. Note that t, and
sion of the cumulative probability distribution of ln Zk : therefore A1,2 , depend on θ implicitly through the Kraus
For x  ln Z typ , Gk (x) has a plateau at unity, and for operators in Eq. 21 [we will sometimes write A1,2 (θ) to
x  ln Z typ it has a plateau at zero, as illustrated in emphasize this dependence].
Fig. 4. The value of the parameter λ is determined in a stan-
As in Ref. [22], we first study the linearized recursion dard way as for Fisher-Kolmogorov-Petrovsky-Piskunov
relation (Eq. 18), and we then take into account the non- waves [51]: so long as the initial condition decays suf-
linear terms. We are interested in the distribution of Zk ficiently fast at large x (which we confirm below), the
9

solution λ that is selected is the one for which the veloc- by symmetry (the two bottom legs of t are equivalent on
ity vθ (λ) is minimal. The “dispersion relation” vθ (λ) is a average).
convex function of λ, and for each fixed θ there is a value Equation 37, together with Eqs. 23 and 24, defines
of λ = λ∗ that gives the minimal velocity: the critical measurement strength θc . Equation 37 is
easy to solve numerically, since it involves only a finite-
∂vθ (λ) dimensional integral over the random parameters of a
= 0, vθ ≡ vθ (λ∗ ). (30)
∂λ λ∗ single node, but in fact it can also be solved analytically.
So far, the identities above relied only on the unitary
If we plug in Eq. 29, we get
invariance property discussed around Eq. 35, and did not
1 ∗ ∗

1 hAλ ln A1 + Aλ2 ln A2 i

depend on the more detailed structure of the node tensor
− ∗2
ln(hAλ1 + Aλ2 i) + ∗ 1 ∗ ∗ = 0. t. For the present model, where t is given by Eq. 12,
λ λ hAλ1 + Aλ2 i
averages like hAλ i can be written as explicit functions
(31)
of θ by sequentially integrating out the various Haar-
For a given value of θ, the above equation fixes λ∗ [and
random objects appearing in the node. This calculation
therefore vθ ≡ vθ (λ∗ )]. At the critical point we have a
is presented in Appendix C and gives
second equation, namely
λ
λ
A1 = A2 =
vθc = 0. (32)
3 × 41−λ (sin 2θ)2λ (cos 2θ)−1 (sin θ)2−4λ − (cos θ)2−4λ

Notice that the first term on the left-hand side of Eq. 31 .
(λ + 1)(λ + 2)(2λ − 1)(λ − 2)(λ − 3)
is equal vθ /λ, which vanishes at θc . Therefore the two (38)
equations which hold at the critical point, Eqs. 31 and
32, reduce to Taking the limit λ = 1/2 in this formula, Eq. 37 becomes
∗ ∗
hAλ1 ln A1 i + hAλ2 ln A2 i = 0. (33) ln γ 75
= , for γ = tan θc , (39)
∗ ∗ γ − 1/γ 256
hAλ1 i + hAλ2 i = 1. (34)
which has the solution
These two equations determine both unknowns, λ∗ and
θc , i.e. they determine the location of the critical point. θc ' 1.42010054727632. (40)
Surprisingly, Eqs. (33) and (34) can be solved analyt- We also obtained θc by evaluating the averages in
ically. Equation 33 can be solved for λ∗ using only an Eq. 37 directly, by numerically integrating over the Haar-
invariance property of the node tensor distribution (we random elements in the node, and we obtained the com-
consider this solution first), while for Eq. (34) we must patible result θc = 1.4201 ± 0.0001.
consider the particular structure of the node tensor. Above we assumed that the travelling wave converged
Since both the weak measurement bases and the uni- to the solution with with λ = λ∗ . For completeness, let
tary operator U are chosen Haar-randomly, the t-tensor us briefly recall the justification for this assumption.
is statistically invariant under unitary rotations on any The solution chosen at late times depends on which
single index. For instance, consider the mapping decays faster with x at large x: the initial con-
tacd → ucc0 tac0 d ; (35) dition 1∗ − G0 (x), or the minimal-velocity wavefront,
1 − G(λ ) (x). For the initial condition, expanding the
the two tensors t and t0 = ut have equal probability. generating function at large x gives
Given this condition,7 the coefficients A1 and A2 satisfy
G0 (x) ' 1 − hZ0 ie−x for x → ∞. (41)
the identity [22] (irrespective of the value of θ)
1/2
Thus, the initial condition effectively corresponds to a
hAi ln Ai i = 0. (36) value of λ = 1 [see Eq. 28]. The selected value of λ at
late times is the minimum between λ∗ and 1. Borrowing
This may be shown by averaging over a single-leg unitary terminology from a closely related problem in directed
rotation like that in Eq. 35 (an analogous argument is polymers [50], we say that the system is in the glass
discussed in App. B). Given this identity, Eq. 33 implies class in cases where λ converges to λ∗ < 1, while cases
that λ∗ = 1/2 at the critical point, and therefore that the where λ = 1 correspond to the paramagnetic class. In
critical point θc is the value of θ for which (Eq. 34) the present case — the near-critical FMPT – we have
λ∗ < 1, as discussed above, so the solution with λ∗ is the
hA1 (θc )1/2 i + hA2 (θc )1/2 i = 1, (37)
relevant one (see Ref. [22] for further discussion).
where we have made the dependence of Ai on θ explicit.
In the present model the two terms in Eq. 37 are equal
B. Scaling behavior near the critical point

We now consider scaling behavior of Zk near θc . We


7 Here we also use the fact that Ai > 0 with probability 1. focus on two specific quantities: the size of the order
10

≡ limk→∞ Zktyp in the entangling phase θ


typ
parameter Z∞
as one closely approaches the transition (θ < θc ); and 0
the dependence of Zktyp on k exactly at the critical point, 1.5
θ = θc . 1.4
Once we leave the disentangled phase, the linear re- −10
cursion relation discussed above is no longer sufficient: 1.3

ln Zktyp
the nonlinear terms in Eq. 18 are essential to prevent 1.2
Z from diverging. However, previous work [22] showed −20
that ultimately the critical scaling properties can be ex- 1.1
θc = 1.42
pressed in terms of the velocity function vθ (λ) (Eq. 29).
−30 1.0
Here we briefly list the main results. (We will review the
argument for these scaling forms in Sec. IV and App. E.) 0.9
Just on the entangling side of the critical point, −40 0.8
  0 50 100 150
typ π k
Z∞ ∼ exp − . (42)
|Im λθ |
(λ) FIG. 5. The evolution of Z typ with the tree depth k for various
Here λθ is the solution of the equation vθ = 0. Expand- θ in the forced measurement case. The theoretical prediction
ing Eq. 29 around the critical point, we find for the critical measurement rate, θc = 1.42, is indicated by
  the dashed line. When the measurements are weak enough
typ C that θ < θc , Zktyp converges to a nonzero value in the limit
Z∞ ∼ exp − √ for θc − θ  1, (43)
θc − θ k → ∞. When the measurements are stronger, θ ≥ θc , Zktyp
tends to zero as k → ∞, so that ln Z typ tends to −∞.
where C is a constant coefficient. Using Eq. 29 and
Eq. 42, we can calculate the coefficient C,
q
P 1/2 because of the exponential growth of the number of nodes
i hAi (θc )(ln Ai (θc )) i
2
π with k. We overcome this limitation by using the pool
C=√ r . (44)
2 method described in Refs. [52–54]. Briefly, this method

P 1/2
i h∂θ Ai (θ)|θc i

Npool
involves storing a large set (“pool”) {Zka }a=1 of values
of Zk corresponding to trees of finite size k. This pool
Numerically evaluating these averages gives C = 2.903 ± is a way of approximately capturing the probability dis-
0.001. Taking appropriate derivatives of Eq. 38 (with tribution µk of Zk , via a large set of values drawn from
respect to λ to obtain the numerator of Eq. 44 and with this distribution, with larger values of the pool size Npool
respect to θ to obtain the denominator) gives: giving a better approximation. A pool {Zk+1 } corre-
sponding to trees of size k + 1 can then be produced by
C ' 2.90330810201297. (45)
randomly combining the values from the pool {Zk } using
Exactly at the critical point, θ = θc , the velocity vθ Eq. 18. Our numerical averages for ln Zktyp are taken over
vanishes and the value of ln Zktyp evolves sub-ballistically all elements of the pool. All our results in the main text
with k. Reference [22] argued that were obtained with a pool size 106 , and we show in Ap-
pendix D that the results we present are well-converged
ln Zktyp ∼ −k 1/3 . (46) as a function of the pool size.
Figure 5 shows numerical results for ln Zktyp as a func-
Finally, we note that λ∗ = 1/2 can be viewed as a criti- tion of k for different values of the measurement strength
cal exponent determining the breadth of the probability θ. The existence of a critical measurement strength
distribution of Zk near criticality [22]. We discuss this θc ≈ 1.42 can be clearly seen as a demarcation between
further in Sec. IV D. two different classes of curves ln Zktyp : those that satu-
Equations 40, 43 and 46 are our main predictions for rate to a finite value at k → ∞ and those that continue
the critical properties of the FMPT. toward −∞ (corresponding to Zk → 0).
Figure 6 shows numerical results for the scaling of the
typ
C. Numerical results saturation value ln Zk→∞ on the entangling side of the
transition. The red dashed line indicates the theoretical
prediction of Eq. 43.
In order to check the above predictions for θc and for
the critical scaling, we perform numerical simulations of Figure 7 shows the behavior of Zktyp at the
the collapse process. theoretically-predicted critical point, θc = 1.42. The red
A direct approach, constructing many realizations of dashed line shows the theoretical prediction of Eq. 46,
large trees, would be prohibitively expensive for large k ln Zktyp (θ = θc ) ∼ −k 1/3 .
11

sociated with a given node of the tree are no longer fixed,


but are instead randomly chosen with probabilities deter-
3 mined by the Born rule (leading to additional quantities
to average over in each node). Since the outcomes of dif-
ln(− ln Z∞typ)

ferent measurements can be correlated, the local three-


index tensors t = t(σ1 , σ2 , σ20 ) that make up the network
2 are no longer independently random objects: they now
have nontrivial correlations not only within a node but
also between different nodes.
We show in this section that averaging over the Born
1 rule outcomes leads to different theoretical results for the
Eq. (43)
MPT as compared to the FMPT. Again we find that a
travelling wave picture can be developed. Notably, how-
−4 −3 −2 −1 ever, while the critical FMPT mapped to a travelling
ln(θc − θ ) wave problem in the “glass” class (with λ∗ = 1/2), the
critical MPT maps to the case λ∗ = 1 which is at the
FIG. 6. The scaling behavior of Zk→∞typ
, as θ approaches the boundary of the glass class with the paramagnetic class.
critical point from the entangling phase, is plotted for the case The MPT still allows for two distinct dynamical phases,
of forced measurements. The blue points show results from entangling and disentangling, separated by a critical mea-
our numeric simulation, and the red dashed line shows the surement strength θ = θc that is distinct from that of the
dependence predicted by Eq. 43 for the limit of asymptotically FMPT case. The scaling of Zktyp as a function of k and
small θc − θ. θ is qualitatively similar to that in the MPT, but there
are universal differences between the problems that are
reflected in the probability distribution of Zk .
40
30
A. Linearized recursion relation
20
− ln Zktyp

We wish to consider how the stochastic process de-


fined by the node t transforms the probability distribu-
10 tion µk (Z) into the probability distribution µk+1 (Z).
For the recursive step, we consider (as in the previous
section) a node “event” in which we start with known “in-
∼ k1/3 put” density matrices ρk,1 and ρk,2 for the two spins. By
5
randomly choosing the necessary Haar-random unitaries,
1 10 100 and choosing the measurement outcomes with the Born-
k rule probabilities, we obtain a random “output” density
matrix ρk+1 .
FIG. 7. The scaling behavior of ln Zktyp exactly at the critical As in the previous section, it is sufficient to consider
point is plotted for the case of forced measurements. Numer- the case where the input density matrices ρk,1 and ρk,2
ical results are shown as blue dots. For comparison, the red have the diagonal form in Eq. 16. This simplification is
dashed line shows the theoretical prediction − ln Zktyp ∼ k1/3 , possible because the probability distribution of the out-
which is a straight line in the log-log plot with an undeter- put density matrix, given the inputs, depends only on
mined vertical offset. the eigenvalues of the inputs, and not on their eigenvec-
tors. This irrelevance of the eigenvectors follows from
the Haar-randomness of the unitaries in the node, which
IV. REAL MEASUREMENTS
means that there is no “preferred basis” for the spins.
(For a formal argument, see App. A.)
The preceding section considered the quantum tree Adopting this diagonal form, the input density matri-
with forced measurements. We found a FMPT with prop- ces are evolved using the t-tensor that defines the op-
erties that are consistent with those found in Ref. [22], eration of a single node, which was given in Sec. II B.
although the details of the tree are slightly different. An Recalling these definitions,
interesting question, which is not directly considered in
Ref. [22], is whether we can use a similar approach to h iaσ20
also reveal the properties of the transition induced by [t(σ1 , σ2 , σ20 )]acd = U (Kσ(1)
1
⊗ K (2)
σ2 ) , (47)
cd
real measurements.
The key difference with the FMPT is that for real mea- This tensor t = t(σ1 , σ2 , σ20 ) depends on the 2-site uni-
surements the measurement outcomes σ1 , σ2 and σ20 as- tary and on the random bases for the measurements (this
12

dependence will be left implicit), and also on the mea- Using the expressions for t in Eqs. 11, 12, and the prop-
surement outcomes (these will sometimes be shown as erty of the Kraus operators
explicit arguments). As before, the output density ma- X
trix is given by Kσ† Kσ = 1, (54)
σ
t(ρ1 ⊗ ρ2 )t†
ρf = . (48) one can check
tr t(ρ1 ⊗ ρ2 )t† P that the above probability p is normalized,
such that σ1 ,σ2 ,σ0 p(σ1 , σ2 , σ20 ) = 1.
2

The measurement outcomes are chosen probabilistically Using Eqs. 51-52, the generating function Gk (x) has
with the Born rule: for given initial states and unitaries, the recursion relation
the probability of a set of outcomes (σ1 , σ2 , σ20 ) may be X
written Gk+1 (x) = hp(s)Gk (x−ln A1 (s, θ))Gk (x−ln A2 (s, θ))i.
s

p(σ1 , σ2 , σ20 ) = tr t(σ1 , σ2 , σ20 )(ρ1 ⊗ ρ2 )t(σ1 , σ2 , σ20 )† , (55)


(49) Here we use s to represent the set of all three measure-
as discussed in Sec. II A. Below we also write this proba- ment outcomes. The angle brackets are the average over
bility as p(s), using s = (σ1 , σ2 , σ20 ) to denote the collec- the Haar-random unitary operations: the two-site ran-
tion of measurement outcomes for the node. For nota- dom unitary U , and the 1-site random unitaries in K (i)
tional simplicity we have suppressed the dependence of which set the bases for the weak measurements. The
p(s) on the initial states and on the random unitaries. argument θ is the strength of weak measurements.
As before, at late times Gk (x) converges to a traveling
Formally, the value of Zk+1 is a function of the initial
wave solution, parameterized by a variable λ which must
states Zk,1 , Zk,2 and of the t-tensor (and through it the
be determined. The solution for a given λ has velocity
measurement outcomes):
1 X

p(s)(Aλ1 (s, θ) + Aλ2 (s, θ)) .



Zk+1 = f t(σ1 , σ2 , σ20 ), Zk1 , Zk2 . vθ (λ) = ln (56)

(50)
λ s
For a fixed set of measurement outcomes, the expansion
For a given θ, there is a special value λ = λ∗ that gives
of the function f in terms of the arguments Zk,1 , Zk,2
the minimal velocity,
proceeds identically to the forced measurement case
(Sec. III A), and thus Eqs. 23 and 24 for the coefficients
∂vθ (λ)
A1 and A2 are still valid for small Zk,1 , Zk,2 : = 0. (57)
∂λ λ=λ∗
Zk+1 ' A1 (σ1 , σ2 , σ20 )Zk,1 + A2 (σ1 , σ2 , σ20 )Zk,2 , (51) The velocity selected is given by
with 
vθ (λ∗ ) if λ∗ ≤ 1
vθ = , (58)
vθ (1) if λ∗ ≥ 1
t11 t21 − t121 t211 2 t11 t12 − t112 t211 2
1 2 1 2
A1 = 2 , A2 = 2 , (52)
(|t111 |2 + |t211 |2 ) (|t111 |2 + |t211 |2 ) and the critical point θc is defined by the vanishing of
this velocity. Recall the physical meaning of this ve-
where the coefficients A1 = A1 (σ1 , σ2 , σ20 ) and locity: in the disentangling phase, vθ is negative, and
A2 = A2 (σ1 , σ2 , σ20 ) are now dependent on the measure- Zktyp ∼ e−|vθ |k .
ment outcomes. If λ∗ ≤ 1, the critical point is given by solving the
This dependence on the measurement outcomes means following equations for the unknowns λ∗ and θc :
that, unlike the forced measurement case, the coefficients
Ai in Eq. 51 are not chosen independently of the Zk,i , vθc (λ∗ ) = 0, ∂λ∗ vθc (λ∗ ) = 0. (59)
since the probability of given measurement outcomes is
itself dependent on Zk,1 and Zk,2 . Fortunately, however, These equations may be rewritten as
do not need to consider the full dependence. In partic- X X D ∗
E
ular, in the disentangling phase, where Zk becomes ar- p(s)Ai (s, θc )λ = 1, (60)
bitrarily small at large k, we can approximate the mea- s i=1,2

surement probabilities by their Z → 0 limits. This ap-


X X D ∗
E
p(s)Ai (s, θc )λ ln Ai (s, θc ) = 0. (61)
proximation is sufficient to locate the transition. s i=1,2
Expanding Eq. 49 in Zk,1 , Zk,2 and retaining only the
zeroth-order term, and using the fact that ρ1,2 are diag- In Appendix B we show that the coefficients A1 (s, θ),
onal in our chosen basis and have the form in Eq. 16, we A2 (s, θ) satisfy the striking identity
find X
hp(s)Ai (s, θ) ln Ai (s, θ)i = 0 (62)
p(σ1 , σ2 , σ20 ) ' |t111 |2 + |t211 |2 . (53) s
13

for both i = 1 and i = 2. As a result, Eq. 61 is satisfied detailed and rigorous study of the recursion relations to
by λ∗ = 1. Equation 60 then gives the value of θc : the future.
X X Slightly inside the weak-measurement phase, we find
hp(s)Ai (s, θc )i = 1. (63) that
s i=1,2
C0
 
typ
This equation may also be written Z∞ ∼ exp − √ (for θc − θ  1) (67)
θc − θ
16 hp(↑, ↑, ↑)A1 ((↑, ↑, ↑), θc )i = 1, (64) where the constant C 0 is discussed below. Exactly at the
since the average in Eq. 63 is independent of s (which critical point,
runs over 8 possible measurement sequences) and of i. ln Zktyp ∼ −k 1/3 (for θ = θc ). (68)
Thus, at the critical point the wavefront solution has
λ∗ = 1, which means that the linear recursion relation is These scaling forms are compared with simulation data
poised exactly at the boundary point between the glass in Sec. IV C. They are similar to those for forced mea-
and paramagnetic classes (which yield different critical surements (Sec. III). On the other hand, λ∗ can be viewed
scaling once nonlinearities are included [22]). Note that as a critical exponent (governing the breadth of the criti-
λ∗ = 1 only at the critical point, and in general the value cal probability distribution for Zk ) that distinguishes the
of λ∗ varies with θ near the critical point as λ∗ −1 ∝ θc −θ. real and forced measurement critical points. This differ-
Numerically integrating the left-hand side of Eq. 63 ence is discussed in Sec. IV D.
allows us to estimate the critical measurement strength Now we discuss the order parameter scaling in slightly
θc . This process gives θc = 1.100 ± 0.001. more detail. (Some readers may prefer to skip the re-
The above procedure, and the critical point equation mainder of this subsection.)
(63), applies for a larger class of tree models with mea- It is convenient to define
surements, in which the averages h. . .i have the unitary
invariance property discussed in Sec. III A. Surprisingly, H(x) = lim 1 − Gk (x), (69)
k→∞
the simple structure of the node tensor t in the particular
model under study allows the integral to be performed which is nontrivial in the entangling phase. In App. E
analytically. Any average of the form hpκ Aλi i can be eval- we give an exact linear equation for H(x) that holds for
uated explicitly as a function of θ (App. C), in particular x  ln Ztyp , i.e. far to the right of the front. This equa-
tion simplifies further when x is also sufficiently large and
hp(s)Aλi i =
negative. In this intermediate regime (loosely speaking,
3 × 21−2λ (sin 2θ)2λ (cos 2θ)−1 (sin θ)4−4λ − (cos θ)4−4λ

negative x with 1  |x|  | ln Z typ |), the analysis is
(λ + 1)(λ + 2)(λ − 3)(λ − 4)(2λ − 2) similar to that in the previous section, where we may
take H(x) ∼ e−λx . For the critical scaling on the entan-
The critical point equation (64) then becomes gling side of the transition, however, we are looking for
a solution that is stationary (independent of k), so the
ln γ 3
= , for γ = tan θc , (65) parameter λ must be chosen such that the traveling wave
γ 2 − 1/γ 2 16 velocity vθ (λ) vanishes. The key point is that satisfying
which has the solution this condition requires a complex λ [55]:
p
θc ' 1.10010302468401. (66) λ = 1 ± ic θc − θ + O(θc − θ). (70)

(Here c is a constant determined by Eq. 60.) Taking a


B. Scaling behavior near the critical point real combination of the two complex solutions then gives
 p 
The primary qualitative difference between the real H(x) ∝ e−x sin φ + c θc − θ x , (71)
measurement and forced measurement critical points is
that the real measurement case has λ∗ = 1, while the where φ is an undetermined constant.
forced measurement case has λ∗ = 1/2. Recursion rela- By considering how this solution in the intermediate
tions with λ∗ < 1 and λ∗ > 1 generally exhibit different regime 1  |x|  | ln Z typ | matches onto solutions to the
order parameter scaling, so the scaling for the MPT in left and right of this interval we may fix the scaling of
our model is a subtle question.8 Below we give a heuris- Z typ . Note that Eq. 71 gives a logarithmic “slope” that
tic argument fixing the leading scaling, leaving a more is close to −1 for most values of x:

c θc − θ
∂x ln H = −1 + √ (72)
tan(φ + c θc − θ x)
8 The fact that the distributions of the Ai in Eq. 51 are depen-
dent on the Zk is also a new feature compared to the previously The exceptions are close to the zeroes of the tangent, for
studied cases. which the second term is no longer small.
14

As x approaches the left hand side of the regime where θ


(72) is valid, i.e. as x approaches ln Z typ , we expect that 0 1.25
the magnitude |∂x ln H| of the slope decreases towards 1.20
zero, in order to match the plateau in H for x  ln Z typ .
This condition suggests that the argument of the tan −5
1.10
function should vanish at a value of x close to ln Z typ

ln Zktyp
[22]. Consequently,
−10 1.00
typ C0 θc = 1.10
ln Z ∼ −√ , (73)
θc − θ −15 0.90
0
with C = φ/c. In principle, we should now use the
matching condition on the other side of the intermedi- −20 0.80
ate regime (where x approaches −1) to fix the numerical 0 50 100 150 200 250
value of φ. We suspect that the value φ = π/2 is required, k
in order for Eq. 72 to be consistent with an expansion of
the generating function in terms of moments of Zk (see FIG. 8. The evolution of Z typ with the depth k of the tree for
App. E). However, confirming this would require a more the case of real measurements, plotted for different values of
careful analysis of corrections to Eq. 72. the measurement strength θ. The red dashed line represents
Finally, given Eq. 71, the form (68) for the k- the value of θ corresponding to the theoretically predicted
dependence of Zktyp exactly at the critical point follows critical point (θc ≈ 1.10). The entangling phase (θ < θc ) is
typ
from the same heuristic argument as in the forced mea- characterized by a nonzero value of Zk→∞ , and the disentan-
typ typ
surement case.9 gling phase (θ > θc ) has Zk→∞ = 0, i.e. ln Zk→∞ = −∞.

C. Numerical results
2.0
Figure 8 shows the evolution of ln Zktyp
with the tree
depth k for different values of the measurement strength 1.5
ln Z∞typ

θ. As in the forced measurement case, one can see two


phases separated by the critical measurement strength 1.0
θc . In the entangling phase, θ < θc , the value of ln Zktyp
is finite in the limit k → ∞, so that an initially mixed 0.5
state remains mixed regardless of the size of the tree. On
the other hand, in the disentangling phase, θ > θc , the 0.0 Eq. (67)
value of ln Zktyp is −∞ in the limit k → ∞, which means
that an initially mixed state is completely purified in the −4 −3 −2
limit of a large tree. The red curve in the plot represents ln(θc − θ )
the critical value θc obtained by the linear recursion re-
lation; one can see that this value is consistent with the
typ
numerical simulation. FIG. 9. The scaling behavior of Zk→∞ as θ approaches the
typ critical value from the entangling side of the transition is plot-
Figure 9 shows the scaling behavior of ln Zk→∞ as one
approaches the transition from the entangling side. The ted for the case of real measurements. The blue points show
results from our numeric simulation and the red dashed line
red dashed line shows the predicted dependence given by
shows the functional dependence predicted by Eq. 67 for the
Eq. 67. The predicted scaling of ln Zktyp with k at θ = θc limit of asymptotically small θc − θ. The value of the param-
is also confirmed in Fig. 10, which shows ln Zktyp ∼ −k 1/3 . eter C 0 is calculated with φ = π/2.

D. Comparison: real vs. forced measurements dynamical phases as a function of increasing measure-
ment strength. However, the real measurement case has
In closing this section we briefly comment on the dif- a smaller critical value of θ, which means that the en-
ferences between the forced measurement and real mea- tangling phase is smaller for real measurements than for
surement cases. In both cases the system exhibits a forced measurements.
phase transition between entangling and disentangling This difference between the two cases implies that real
measurements are more strongly purifying than forced
measurements for the present dynamics. To get some
intuition for the difference between forced and real mea-
9 See Sec. IV.H.4 of Ref. [22]. surements in terms of purification, one can consider the
15

20 (a)
λ = 0.5
10 −2.5
− ln Zktyp

lnhZik→∞
5 −5.0

−7.5

∼ k1/3 −10.0
1
1 10 100
k −20 −15 −10 −5 0
typ
ln Zk→∞
FIG. 10. The value of ln Zktyp at the critical point (b)
θ = θc ≈ 1.10 is plotted for the case of real measurements.
0.0 λ = 1.0
The red dashed line shows the dependence ln Zktyp ∼ −k1/3
predicted by Eq. 68.
−2.5

lnhZik→∞
simple example of a single weak measurement on a sin- −5.0
gle spin. Specifically, we compare the effect of a true
measurement to the effect of a forced measurement in −7.5
which the outcome is chosen independently of the state
and with 50% probability for ↑ or ↓. −10.0
Let us visualize the spin’s initial density matrix ρinit
as a point in the interior of the Bloch sphere (recall that −12 −8 −4 0
pure states lie on the surface of the Bloch sphere, while typ
ln Zk→∞
mixed states lie in the interior). Without loss of gener-
ality, assume that this initial density matrix is closer to
the north pole (|↑i) than to the south pole (|↓i). FIG. 11. The relationship between the typical value of Z
Now make a very weak [56] measurement (θ = π/4 + ) (defined by Eq. 17) and the average value of Z depends on the
in the Sz basis. After the measurement, the density ma- parameter λ. The expected relationship is lnhZi ∼ λ ln Z typ ,
trix is altered so that the corresponding point in the with λ = 1/2 for forced measurements and λ = 1 for true
measurements. Plot (a) compares this prediction (red dashed
Bloch sphere is displaced very slightly. If the measure-
line) with data (blue dots) for forced measurements. Plot (b)
ment outcome is ↑, the displacement takes the point shows the case of true measurements. The somewhat worse
closer to the north pole. Since the state is initially closer agreement for case (b) is likely due to a logarithmic prefactor
to the north pole, this displacement also brings it closer that we have neglected in the relation between Z typ and hZi.
to the surface of the Bloch sphere, i.e. it is purifying. By
contrast, if the measurement outcome is ↓, the density
matrix gets closer to the south pole and becomes less that the nodes of the tensor network T are no longer inde-
pure. pendent. The measurement outcomes at a node, which
Real measurements make the first outcome (the puri- go into determining the node tensor t(σ1 , σ2 , σ20 ), have
fying one) more likely, while for forced measurements the probabilities that depend on the density matrices ρ1 , ρ2
two outcomes are equally likely. The result is that, after coming from the earlier part of the tree.10
averaging over outcomes, the real measurement gives a Nonetheless, the Haar-randomness of the 2-site uni-
mean increase in purity of order , while the forced mea- taries means that the ensemble of density matrices ρk
surement gives a mean change that is higher-order in  for a tree of k generations can be characterized only by
(we omit the detailed calculation).
The other prominent difference between the MPT and
FMPT is the different value of the critical exponent λ
10 In our treatment of the node we considered (without loss of gener-
that characterizes the two transitions; λ = 1/2 for the
ality) the case where ρ1 and ρ2 are diagonal. The correlations be-
FMPT, which is within the glass class, while λ = 1 for
tween ρi and t then imply that the resulting distribution for the
the MPT, which is at the boundary between the glass and node tensor t is not invariant under the rotations ta a
bc → tb0 c ub0 b
paramagnetic classes. Ultimately this difference arises or ta
bc → ta u 0 . That is, ρ
bc0 c c 1,2 have picked out preferred states
because the Born rule probability p(s) depends on the on the bonds, and the distribution of t is sensitive to these pre-
set of measurement outcomes s. This dependence implies ferred states.
16

a distribution of singular values Zk . This allows us to


treat the problem via a recurrence relation for Zk .
While this difference in the value of λ is not reflected
in the critical scaling of ln Zktyp near the transition (com-
pare Eqs. 43 and 46 with Eqs. 67 and 68), it does have
an observable consequence in terms of the distribution
of values of Zk among different realizations of the tensor
network. Specifically, the exponent λ controls the proba-
bility density P (ln Z) d ln Z for ln Z [22, 50]. We consider
the regime just inside the entangling phase, with finite
typ
asymptotic values Z typ = Zk→∞ and hZi = hZk ik→∞
(similar considerations apply just on the other side of the
transition, and at the critical point, for the probability
distribution at large finite k). When Z typ is very small,
there is a wide range for which 1  | ln Z|  | ln Z typ |,
and in this range

P (ln Z) ∝ (Z typ /Z)λ (74)

when λ ≤ 1 (neglecting a possible more slowly-varying


prefactor). Consequently, when Z typ is small the re-
λ
lationship between hZi and Z typ is hZi ∼ (Z typ )
(again, neglecting a possible more slowly-varying pref- FIG. 12. Schematic picture of the proposed expansion pro-
tocol. We use the ∪-shaped connection at the initial time
actor). This relationship is confirmed by our simulation
to represent the initial GHZ state between one system spin
data in Fig. 11 for both the forced measurement and and the reference spin. The world line of the reference spin
real measurement cases. Specifically, the two cases ex- is shown by the red line. A possible choice for the design of
hibit a different slope λ when lnhZk→∞ i is plotted against each node, following Fig. 3, is shown at the bottom.
typ
ln Zk→∞ . This plot provides an independent confirma-
tion of the different value of λ between the two cases.
Thus, formally the expansion process with true mea-
surements is an intermediate case between the two cases
V. THE EXPANSION PROCESS AND that were solved analytically in the previous sections. It
EXPERIMENTAL PROTOCOLS would be interesting to understand how this intermediate
case differs in its critical properties from the other two
Having discussed the collapse process in the previous cases. A naive guess is that it has qualitatively similar
sections, we now turn to the phase transition in the ex- scaling properties, with some value of the critical expo-
pansion process, and we suggest a schematic protocol for nent λ. We leave an investigation of such mixed dynamics
detecting it experimentally. to the future.
The critical properties of the expansion process with Instead, we comment on the implementation of the
forced measurements have been solved in Sec. III, since measurement phase transition in the expansion process
they map to those of the collapse process FMPT (the on a quantum device. Such an experiment does not re-
same ensemble of tensor networks arises in both cases, quire any postselection, in contrast to usual versions of
with opposite choices for the arrow of time). We do the MPT. See Refs. [12, 44–48] for recent approaches
not solve the expansion process with true measurements to the statistical challenge associated with postselection.
here. Unlike for forced measurements, reversing the ar- In particular, Refs. [47, 48] discuss approaches involv-
row of time does not map the expansion process with true ing “hybrid observables” that mix classical computations
measurements to the collapse process with true measure- and quantum measurements. Our proposal below follows
ments. Instead, reversing the arrow of time maps the the spirit of these works, as it makes use of the efficient
expansion process with true (weak11 ) measurements to classical computability (given the measurement record)
a “mixed” version of the collapse process, in which the of certain expectation values. Here, the tree structure
weak measurements are true measurements, but the pro- means that these classical computations are particularly
jective measurements are forced. This equivalence can simple. It would also possible to study the collapse pro-
be shown using the formulas in Sec. II B. cess experimentally: we comment on this in Sec. VI.
The discussion in the remainder of this section is not
restricted to any particular design of the “node” interac-
tion, so we will abstract away from the particular models
11 Note that in the expansion process defined in Sec. II there are we have considered so far, and imagine a more general
only weak measurements, no strong measurements. version of the expansion process. For example, it is not
17

necessary that the interaction unitaries be random: we one could in principle obtain observables like the order
could consider a protocol in which the only randomness parameter hZi directly from this computation. We will
comes from the measurement outcomes.12 denote this estimate as hZiS , since it makes use of mea-
Extrapolating from numerical explorations of the col- surement data for the system S alone (not for the ref-
lapse process, it is likely that clear evidence of the phase erence spin). In reality, it is crucial to make a further
transition could be obtained from an experiment on a measurement of the reference spin in each realization, in
relatively small number of qubits. order to obtain independent confirmation of the result of
We consider an expansion process that starts with the the density matrix reconstruction. A conceptually simple
single initial “system” spin in in a maximally entangled way to do this is as follows.
(GHZ) state with a “reference” spin, denoted R [12]. We In each run one has the recursively computed estimate
run the expansion process for the system, recruiting addi- of the polarization vector ~n and therefore of its normal-
tional spins at each time step. A cartoon of this dynamics ized version n̂. One can now make a Pauli measurement
is illustrated in Fig. 12, where the ∪-shaped connection in the appropriately rotated frame, i.e. a measurement of
at the bottom represents the initial entangled state of the n̂ · ~σ . This measurement yields an outcome τ = ±1, with
“system” and “reference” spins. We will detect the MPT the outcome τ = −1 having a probability Z. Therefore
using measurements of the reference R. these measurements give us an estimate of the average
At the end of the expansion process we have an en- value of the order parameter Z:
tangled state |Ψi of R and the 2t system spins. The
Schmidt values for a √ bipartition
√ of this state between R 1
and the system are { 1 − Z, Z}, where Z is the order hZiSR = (1 − hτ i) , (77)
2
parameter that we have focused on in previous sections.
Tracing out the system, the reduced density matrix for where the averaging is over different runs of the exper-
the reference ρR has eigenvalues 1 − Z and Z. iment. The subscript on hZiSR indicates that this esti-
To begin, imagine an idealized experiment with no un- mate combines measurement data from both the system
controlled noise. In a given run of the experiment we S and the reference R. hZiSR should agree with the esti-
obtain a measurement record mW . The final state ρR of mate hZiS which uses measurement data from the system
the reference will depend on mW . However the recursive only.
structure of the tree means that a very straightforward Equation (77) implies that the MPT for the expan-
processing of the experimental data mW allows us to de- sion process corresponds to a phase transition in the pre-
duce this final state ρR . The computation simply iterates dictability of the final measurement on the reference spin.
a recursion relation for 2×2 density matrices. This recur- In the disentangling/strong measurement phase, the re-
sion has exactly the same structure as Eq. 22: we start duced density matrix ρR for the reference spin is a pure
at the leaves of the tree and iterate towards the trunk (in state in the limit k → ∞ of large tree depth. Con-
the present setting this process corresponds to iterating sequently, the final measurement of the reference spin
backwards in time).13 Notably, the computational effort (the value of τ ) can be predicted with perfect accuracy
is proportional only to the number of nodes in the tree: by someone keeping track of the measurement record of
it is sufficient to work with only 2 × 2 density matrices, the system spins (and using them to perform the recur-
rather than a many-body quantum state. sive calculation of ρR ). On the other hand, in the en-
Let us write ρR in the form tangling/weak measurement phase the reference spin re-
mains in a mixed state even as k → ∞, and consequently
ρR = (1 + ~n · ~σ )/2, (75) the final measurement on the reference spin cannot be
predicted with perfect accuracy.
where ~σ is the vector of Pauli matrices, and The protocol described above requires one to make a
spin measurement in an arbitrary basis defined by n̂,
|~n| = 1 − 2Z. (76)
which might be challenging. However, even if measure-
In the idealised experiment, ρR can be computed ana- ments can only be made in a fixed basis, say σz , the
lytically from the measurement record in each run, and average value of the order parameter, hZi, can still be
extracted. Specifically,
  
1 τ
hZi = 1− , (78)
12 It is also possible to replace weak measurements with projective 2 n̂z
ones, at the cost of involving more spins in the node. (This equiv-
alence is a consequence of the fact that a weak measurement can where nz is the z-component of the vector ~n that is cal-
be effected using an additional ancilla spin that gets projectively
culated theoretically for each run, τ = ±1 is the σz mea-
measured.)
13 The starting state for this recursion corresponds to maximally surement outcome for the reference, and again the av-
mixed states for each of the “leaves” of the tree. For each node, erage is over runs. The value of hZi calculated in this
the node tensor t used for the recursive step depends on the way using the experimental measurements of σz should
measurement outcomes that were obtained at that node. exhibit the critical scaling of the MPT.
18

VI. OUTLOOK sible disorder, or can other critical models realize broader
order parameter distributions?
In this paper we have defined and explored the
measurement-induced entanglement phase transition in It is worth pointing out that our model contains no
dynamical quantum trees. We have adopted the perspec- randomness in the time or location of measurements (in
tive of purification [7], so that the entanglement transi- the FMPT case there is no randomness in the measure-
tion is characterized by a single “order parameter” Zk ment outcomes either). Nonetheless, a phase transition
that describes the purity of the final state after a k-step still exists as a function of the measurement strength.
protocol of measurement and unitary evolution. Forthcoming work shows that a completely uniform tree
tensor network, with no randomness of any kind, can also
Our primary contribution is to extend the approach
show an entanglement transition [34]. The model studied
of Ref. [22] in order to provide analytical solutions for a
here, and extensions of it, may provide insight into the
transition (for the collapse process) not only for forced
role of different kinds (and strengths) of randomness at
measurements, but also for real measurements. We have
the MPT.
shown that a transition exists in both cases (the FMPT
and the MPT), with distinct nontrivial values of the crit-
An outstanding task that we have not tackled here is
ical measurement strength θc . The value of θc is some-
to develop a recursive approach to the expansion process
what smaller in the real measurement case, which in
with real measurements. Finding an analytical solution
the language of purification arises because real (weak)
to this problem would be interesting. Does this process
measurements are more purifying than forced ones. At
correspond to a phase transition of the same general type
the most basic level, the MPT and FMPT exhibit simi-
as those studied here, with some value of the critical ex-
lar critical behavior near the transition, as captured by
ponent λ?
Eqs. 43–46 and Eqs. 67–68, but there are universal dif-
ferences between the two types of transition. In Sec. V we suggested a schematic experimental im-
The most striking difference between the MPT and plementation of the MPT for the expansion process. It
FMPT is in the values of the critical exponent λ, which would also be possible to realize the collapse process ex-
characterizes the breadth of the probability distribution perimentally. One general approach [47] is to compare
for the order parameter Z (or equivalently for the entan- the evolution of different initial states that are condi-
glement) at the critical point. The MPT has the larger tioned on the same measurement outcomes. In the col-
value λ = 1, as compared with λ = 1/2 for the FMPT. lapse process, it is sufficient to compare the output single-
This difference implies a narrower distribution of Zk val- spin state for distinct initial many-body states. In out-
ues in the MPT case for a given tree size k. line, one initial state is run experimentally and the other
The models are analysed using an auxiliary recursion initial state is simulated classically, enforcing the same
relation for Z. Interestingly, the value λ = 1 lies precisely outcomes. The average distance between the outputs can
at the boundary between two classes of phase transition then be estimated with an appropriate measurement, and
for such recursion relations. The analysis here shows that can serve as an order parameter (in the strong monitor-
the value λ = 1 is protected for a larger class of models, ing phase, the tree tensor network has a single dominant
in which the node operations have a statistical invari- singular value, and so projects almost all inputs onto the
ance under unitary rotations. But it is possible that rel- same output). As in the case of Clifford circuits [47], the
atively small modifications to the quantum tree model initial state that is run experimentally can be a state that
that take it outside this class of models could push λ for is not efficiently representable classically.
the MPT to values larger than unity. In such a model,
the MPT and FMPT would show more drastically differ- Separately, it would also be interesting to under-
ent universal properties. (According to the arguments of stand whether the expansion or collapse processes are
Ref. [22], at λ > 1 the onset of the order parameter near related to any natural quantum information task. One
the transition is as a power law, rather than exponential could also explore quantum versions of classical message-
as in the model studied here.) Studying models in this broadcasting processes with a tree-like structure, with
broader class may therefore be a promising goal. These potential transmission errors on the bonds of the tree.
differences are also interesting in relation to possible field These classical processes can show nontrivial phase tran-
theories for the MPT. sitions in the information communicated from the root
One could similarly ask whether smaller values of the to the leaves [57].
exponent λ are realizable when the statistical unitary in-
variance property of the current models is relaxed. This The results here were obtained using a recursion re-
property enforces λ = 1 for the collapse process with lation for a random variable Zk characterizing the en-
real measurements and λ = 1/2 for the processes with tanglement (or purity) of a given tree. An open ques-
forced measurements (at their respective critical points). tion is how to recover them using the formalism of the
Heuristically, smaller λ corresponds to stronger disorder effective “lattice magnet” that can be obtained using
and a broader distribution of entanglement. Do the Haar- the replica trick by integrating out all the random uni-
invariant tree ensembles correspond to the strongest pos- taries [8, 9, 22, 27, 29, 33].
19

ACKNOWLEDGMENTS unitary transformation at the base of the tensor network.


That is, regarding T as a 2 × M rectangular matrix (see
The authors are grateful to Sthitadhi Roy and Sec. II B),
Jonathan Ruhman for helpful discussions. Part of this
work was performed at the Aspen Center for Physics, F (T ) = F (T Y ) (A4)
which is supported by National Science Foundation grant
PHY-1607611. for any M × M unitary Y . For example, any function
of the order parameter Z has this property, since the
singular values of T are unchanged by a unitary rotation.
Appendix A: Gauge transformation of tree Now we wish to show that the averages of F are
the same in two different ensembles: first, the ensemble
where the single-site unitaries u and v are Haar-random,
In Sec. II A we stated that it did not matter whether
and second, the ensemble where they are all fixed to the
the measurements were performed in the Sz basis, or in
identity. The latter corresponds to taking all measure-
randomized bases. In this Appendix we show this using
ments in the Sz basis. (The case where the us are random
the expressions in Sec. II B for the probability of a given
while the vs are fixed, or vice versa, can also be shown
tree. We consider the case of true measurements (the
to be equivalent, in a similar way).
forced measurement case is even simpler) for the collapse
First consider the case where the bases are randomized.
process. The same logic applies to the expansion process.
Then expectation values have the schematic form
Abusing notation, in this Appendix (only) we will de-
note a given tree by

hF i = hhF (T (U, u, v, σ))iσ iU u,v (A5)
T = T (U, u, v, σ). (A1)
The unitary averages involve the Haar measure. The σ
Here U denotes not a single unitary but the complete set average has to be the innermost one, and it is done with
of two-spin unitaries appearing in the tree, u denotes the the Born rule probability, which is a function of T . We
set of single-spin unitaries appearing in the Kraus opera- write this schematically as a function of T (see Sec. II B):
tors (which fix the measurement bases for the weak mea-
surements) and v denotes a set of single-spin unitaries p(σ|U, u, v) = H(T (U, u, v, σ)). (A6)
that fix the bases for the strong measurements. Finally
σ is the set of measurement outcomes. We will also use Since we start with a maximally mixed state, the function
U , u etc. for individual unitaries — this should be clear H is expressed in terms of the trace of T T † , so it obeys
from context. property (A4). Therefore, using (A3),
First, note that, because of the structure of the com-
p(σ|U, u, v) = H(T (V (u, v)U W (u, v), 1, 1, σ)) (A7)
plete tensor network for T , it is possible to absorb almost
all of the us and vs into a redefinition of the U s. For = p(σ|V (u, v)U W (u, v), 1, 1). (A8)
example, in the expression K = uKdiag u† for the Kraus
operator on a bond, we can absorb the u into the U above Now return to the average. Let us label the inner average
the bond, and the u† in to the U below. Similarly each by the probability function used for the average over σ:
v can be absorbed into the U from the same node. In DD E E
this process, a given two-site unitary Ur (here r denotes hF i = hF (T (U, u, v, σ))iσ; p(σ|U,u,v) . (A9)
U u,v
a node) transforms as
Inside the average, we use the identity (A3).
Ur −→ Vr (u, v)Ur Wr (u, v). (A2)
D  E  
Where Vr and Wr are some unitaries that depend on hF i = F Te , (A10)
σ; p(σ|U,u,v)
the adjacent us and the adjacent v. (The details do not U u,v
matter.) The only u† s that do not get absorbed in this
process are the ones at the initial time from the lowest where Te stands for Te = T (V (u, v)U W (u, v), 1, 1, σ). We
layer of Ks. If we denote the product of these by X(u), also use the identity for p.
then the above “gauge transformation” gives us the sim- D  E  
ple identity hF i = F T e .
σ; p(σ|V (u,v)U W (u,v),1,1) U u,v
T (U, u, v, σ) = T (V (u, v)U W (u, v), 1, 1, σ) × X(u).
(A3) Now we define U 0 = V (u, v)U W (u, v). We note that for
Again we use a schematic notation where site indices are any fixed u and v, the distribution of U 0 is independent
dropped. of u and v and is Haar.
The final factor, X(u), will drop out because (i) we DD E E
take the initial state to be maximally mixed, and (ii) hF i = hF (T (U 0 , 1, 1, σ))iσ; p(σ|U 0 ,1,1) 0 . (A11)
we consider observables F (T ) that are invariant under a U u,v
20

Nothing inside the outer average depends on u, v so we Appendix B: λ∗ parameter: real measurement case
can drop the average over these quantities:
D E In this appendix we derive an identity, used in
hF i = hF (T (U 0 , 1, 1, σ))iσ; p(σ|U 0 ,1,1) (A12) Sec. IV A, which fixes the value of λ∗ for the true mea-
U0
surement case. We start with Eq. 62,
This is the average in the ensemble where all the mea- X ∗ ∗

surements are in the Sz basis. This establishes the claim. hp(s)(Aλ1 (s) ln A1 (s) + Aλ2 (s) ln A2 (s))i = 0, (B1)
s
Let us also confirm a similar invariance which we in-
voked at the beginning of Sec. IV A. Consider a single where s = (σ1 , σ2 , σ20 ) labels measurement outcomes, and
node of the collapse process, with real measurements, we recall from Sec. IV A and Sec. II B that
with a definite initial state of the form ρ1 ⊗ ρ2 . We check
that the probability distribution for the output density t11 t21 − t121 t211 2
1 2
t11 t12 − t112 t211 2
1 2
matrix, A1 (s) = 2 , A2 (s) = 2 ,
(|t111 |2 + |t211 |2 ) (|t111 |2 + |t211 |2 )
ρ ∝ t[ρ1 ⊗ ρ2 ]t† , (A13) (B2)

is left unchanged by a basis rotation for one of the initial and


spins, e.g.
p(s) = |t111 |2 + |t211 |2 , (B3)

ρ1 → wρ1 w . (A14) and finally
This invariance can be seen by manipulations similar to h iaσ20
the above. The node tensor t, defined in Sec. II B, de- tacd = t(σ1 , σ2 , σ20 )acd = U (Kσ(1)
1
⊗ K (2)
σ2 ) . (B4)
cd
pends on a two-site unitary U , on single-site unitaries
u1 , u2 appearing in the Krauss operators, and on mea- Since the average over measurements has been separated
surement outcomes s = (σ1 , σ2 , σ20 ). Schematically, ρ is a out and written explicitly in terms of p(s) = p(σ1 , σ2 , σ20 ),
function of all these and the initial states. Let us write the angle brackets in Eq. B1 represent only the averages
over the Haar-random unitary U and the random bases
ρ = ρ(s, X), where X = (U, u1 , u2 , ρ1 , ρ2 ). (A15) in the K operators.14 Therefore, these averages are taken
with a distribution that is invariant under unitary trans-
Let us denote the probability of outcomes s, given the formations on any index of the tensor t.
initial states and the unitaries, as p(s|X) (see Eq. 9 for Eq. B1 means that the quantity we need to evaluate
the expression). Let us also define is not hAλi i alone, as the forced measurement case, but
hp(s)Aλi (s)i. Since A1 and A2 are statistically equivalent,
Xw = (U, u1 , u2 , wρ1 w† , ρ2 ), (A16) it suffices to examine the average involving A1 . Using
ew = (U [w ⊗ 1], w† u1 , u2 , ρ1 , ρ2 ). Eqs. B2–B4,
X (A17)
* 2λ +
 t111 t221 − t121 t211
Then the structure of the tensor network for ρ gives the λ 1 2 2 2


p(s)A1 (s) = |t11 | + |t11 | 2λ
following invariances, (|t111 |2 + |t211 |2 )
t111 t221 − t121 t211 2λ
* +
ρ(s, Xw ) = ρ(s, X
ew ), p(s|Xw ) = p(s|X
ew ). (A18) = . (B6)
2λ−1
(|t111 |2 + |t211 |2 )
As a result, the probability distribution of ρ, for given
input states, is unchanged by the rotation ρ1 → wρ1 w† . This can be analyzed by a similar method to Ref. [22].
To see this, consider an arbitrary expectation value in- Let M be the matrix Mab ≡ tab1 , in terms of which
volving ρ, with the initial state (wρ1 w† ) ⊗ ρ2 : * +


λ
|det M |
hF (ρ)i = hhF (ρ(Xw ))is;p(s|Xw ) iU,u1 ,u2 (A19) p(s)A1 (s) = 2λ−1
. (B7)
(|M11 |2 + |M21 |2 )
= hhF (ρ(X
ew ))i e iU,u1 ,u2 .
s;p(s|Xw ) (A20)

Looking at Eq. A17 we see that w and w† can be absorbed


into the Haar-random unitaries U and u1 without any 14 We could write these operators as e.g.
change to their distribution, so that F (ρ) is independent (1)
Kσ = u(1) Kσz u(1)† , (B5)
of w. That is, the probability distribution of the output
ρ depends only on the eigenvalues of ρ1 and not on its where Kz is non-random and diagonal in the z basis, and u(1) is
eigenvectors (and similarly for ρ2 ). a Haar random 2 × 2 unitary.
21

We consider the singular value decomposition of M , by standard means (see e.g. [22]) giving

α1−µ − β 1−µ
X
Mab = waµ ηµ vµb , (B8) Cµ (α, β) = . (C3)
µ=1,2 (1 − µ)(α − β)
where w and v are unitary, and, with probability 1, both To get started we again write Xκ,λ in terms of the
singular values are greater than zero. Then Eq. B6 be- matrix
comes
** + + Mab = tab1 (C4)

λ
η12λ η22λ
p(s)A1 (s) =  2λ−1 . (see Eq. B4 for the definition of t) as
2 2 2 2
η1 |v11 | + η2 |v21 |
v η
* +
(B9) | det M |2λ
Xκ,λ = 2λ−κ
. (C5)
Here the averages are over the singular values η and the (|M11 |2 + |M21 |2 )
unitary matrix v. Because of the invariance property of
the t distribution mentioned above, the v average is sim- Let us look at the structure of M . Suppressing the “↑”
ply a Haar average, which reduces to a uniform average subscripts on the Ks that indicate the measurement out-
of the complex vector va1 over the unit sphere, as in [22]. comes, this matrix has the form
Performing the sphere average (see Eqs. C2, C2 below for (1) (2)
the necessary formula) gives Mab = Uba1
0 c0 K 0 Kc0 1 .
bb (C6)
(2)
* +
η 4−2λ 2λ
η 2 − η 2λ 4−2λ
1 η The final factor, yc0 ≡ Kc0 1 , has a single index, i.e. it is
p(s)Aλ1 (s) = 1 2


. (B10) a random vector:
(2 − 2λ)(η12 − η22 )
η  
(2) sin θ 0
This expression is sufficient to obtain the desired identity, yc = Kc0 1 = uc d
0 0 u† . (C7)
0 cos θ dd0 d0 1
even without performing the η average explicitly. Differ-
entiating with respect to λ and taking the limit λ → 1, Because U is anyway Haar-random (and independent of
we obtain K (2) ), the orientation of this vector does not matter. We
are free to replace yc0 with a vector that points in the
hp(s)A1 ln A1 i = 0, (B11) (1, 0) direction, without changing the distribution of M .
However, we must keep track of the norm of y. Let us
for any outcome. Thus we prove Eq. 62. And the critical call this norm R. Since the leftmost unitary on the RHS
point is given by Eq. 60. of (C7) does not change the norm,
 
sin θ 0
Appendix C: Exact calculation of averages R= ~z . (C8)
0 cos θ

The calculations in the previous appendix required Here, ~z is a uniformly random vector (previously written
only the unitary invariance property of the node aver- as u†d0 1 ). Note that
ages, and did not use the specific details of the t tensor.
But for our model it is possible to calculate these aver- R2 = sin2 θ |z1 |2 + cos2 θ |z2 |2 (C9)
ages exactly. We will evaluate the average
is a random variable of the type appearing in the denom-
Xκ,λ (θ) ≡ p(s)κ A1 (s, θ)λ .


(C1) inator in Eq. C2.
After the simplification above, we have
As above, the average is over the two-site unitary U and
the Kraus operators in the node tensor t. The result is in- Mab = R × Uba1
01K 0 ,
(1)
(C10)
bb
dependent of the measurement outcomes s = (σ1 , σ2 , σ20 ),
so we can take s = (↑, ↑, ↑) without loss of generality. The where all three elements are statistically independent.
analogous expression with A2 in place of A1 is also equal Next let us make a singular value decomposition of U ,
to Xκ,λ . or rather of the sub-block of U that appears in M :
Before examining Eq. C1, we define a basic average  U 
that we will need repeatedly, a1 η1 0
Ub0 1 = (vHw)ab ,
0 H= . (C11)
  0 η2U
1
Cµ (α, β) ≡ . (C2)
(α|z1 |2 + β|z2 |2 )µ ~z Here v and w are Haar-random 2 × 2 unitary matrices.
(We have denoted the singular values by ηµU to distinguish
Here ~z = (z1 , z2 ) is a complex unit vector that is averaged them from the singular values of M itself, which were
uniformly over the sphere. The average can be performed denoted ηµ in Sec. B.) The distribution of the singular
22

values ηµU is provided in Ref. [58]. Defining si = (ηiU )2 , We have


we have 0 ≤ si ≤ 1 and 2
|M11 |2 + |M21 |2 = R2 R02 |H x̂| , (C21)
2
P (s1 , s2 ) ds1 ds2 = 6(s1 − s2 ) ds1 ds2 . (C12)
giving
The remaining factor in the definition of Mab involves
|M11 |2 + |M21 |2 = R2 R02 s1 |x̂1 |2 + s2 |x̂2 |2 .

the matrix K (1) , which is: (C22)

All three factors on the right hand side of (C22) are ran-
 
0 sin θ 0
(1)
K =u u0† , (C13) dom variables of precisely the kind appearing in Eq. C2,
0 cos θ
with (a, b) = (sin2 θ, cos2 θ) for the first two factors, and
where u0 is yet another 2×2 Haar-random unitary matrix. (a, b) = (s1 , s2 ) for the third factor.
Altogether, from (C10), (C11) and (C13) we have Combining Eq. C15 and Eq. C22,
 2λ
sin 2θ
 
sin θ 0

02κ−4λ
M = R × v H w u0 u0† . (C14) Xκ,λ = R ~z R z0
0 cos θ 2 ~
** + +
(s1 s2 )λ
Now consider the numerator and denominator in (C5). × (C23)
2λ−κ
Since det AB = det A det B, and the unitary matrices (s1 |x̂1 |2 + s2 |x̂2 |2 ) ~
x s 1 ,s2
have unimodular determinant, we have for the determi-
nant of M Using the definition (C2),
1 2 U U  2λ
| det M | = R η1 η2 sin 2θ. (C15) sin 2θ
2 Xκ,λ = C−κ (s2 θ, c2 θ)C2λ−κ (s2 θ, c2 θ)
2
From Eq. C14 we see that the expression |M11 |2 + |M21 |2 × (s1 s2 )λ C2λ−κ (s1 , s2 ) s1 ,s2 ,


is the norm-squared of the vector
(C24)
   
sin θ 0
R × H w̃ ~z 0 , (C16) where (s2 θ, c2 θ) = (sin2 θ, cos2 θ). Finally we must inte-
0 cos θ
grate over s1 and s2 using (C12),
where we have defined za0 ≡ u0† 0
a1 , and the product wu has (s1 s2 )λ C2λ−κ (s1 , s2 ) s1 ,s2 =


been written as w̃ — which is itself Haar-random — and
finally the unitary rotation v in (C14) has been dropped 12
. (C25)
since it does not change the norm. (1 + λ)(2 + λ)(2 + κ − λ)(3 + κ − λ)
  2 Altogether this gives
sin θ 0
|M11 |2 + |M21 |2 = R2 H w̃ ~z 0 . (C17)

0 cos θ Xκ,λ =
0
The unit vector ~z is again uniformly random. 3 × 41−λ (sin 2θ)2λ (cos 2θ)−2
The vector whose norm is taken in (C17) is obtained (1 + κ)(1 + λ)(2 + λ)(1 + κ − 2λ)(2 + κ − λ)(3 + κ − λ)
by applying the diagonal matrix H of singular values to × (sin θ)2+2κ − (cos θ)2+2κ

the complex vector ~x,
× (sin θ)2+2κ−4λ − (cos θ)2+2κ−4λ .

(C26)
 
sin θ 0
~x ≡ w̃ ~z 0 . (C18) The two cases of interest to us here are X0,λ (for the
0 cos θ
FMPT) and X1,λ (for the MPT).
This vector can be written
X0,λ =
0
~x = R x̂. (C19) −3 × 41−λ (sin 2θ)2λ (cos 2θ)−1 (sin θ)2−4λ − (cos θ)2−4λ


(1 + λ)(2 + λ)(1 − 2λ)(2 − λ)(3 − λ)


Here, the norm R0 is given by
(C27)
R02 = sin2 θ |z10 |2 + cos2 θ |z20 |2 , (C20) and
which is another random variable distributed in precisely X1,λ =
the same way as R2 above (meaning that we will be able
−3 × 21−2λ (sin 2θ)2λ (cos 2θ)−1 (sin θ)4−4λ − (cos θ)4−4λ

to use Eq. C3). The unit vector x̂ in (C19) is uniformly
distributed on the sphere, because of the Haar-random (1 + λ)(2 + λ)(3 − λ)(4 − λ)(2 − 2λ)
unitary w̃ in (C18). (C28)
23

Some simple checks of these formulas are the values (a)


X0,0 = 1, X1,0 = 1/8, X0,1 = 1. (C29)

The second of these is the inverse of the number of pos-


sible measurement
P outcomes and follows from the nor- −20
malisation s p(s) = 1, and the third is a general result

ln Zktyp
from the unitary invariance property.
The values that we need to locate the critical points N = 102
are X0,1/2 for the FMPT and X1,1 for the MPT. Defining N = 102.5
−40 N = 103
N = 103.5
γ ≡ tan θ, (C30) N = 104
N = 104.5
these values are N = 105

128 ln γ 1 ln γ −60
X0,1/2 = , X1,1 = . (C31) 0 200 400 600 800
75 γ − 1/γ 3 γ 2 − 1/γ 2 k
For the FMPT we have the critical point equation (b)
1
X0,1/2 = (C32)
2
−5
which gives

θcFMPT ' 1.42010054727632. (C33) ln Zktyp −10


N = 102
For the MPT the critical point equation is N = 102.5
−15 N = 103
1 N = 103.5
X1,1 = (C34)
16 N = 104
N = 104.5
which gives −20 N = 105

θcMPT ' 1.10010302468401. (C35) 0 200 400 600 800


k
Appendix D: Finite pool-size effect
FIG. 13. Convergence of results with increase of pool size.
(a) The forced measurement case with ln(θc − θ) = −3.5. (b)
In this work we use the pool method to approximate The real measurement case with ln(θc − θ) = −3.9. Here the
the recursive evolution of the probability distribution of relative error between the two curves for the largest pool sizes
the density matrix in the quantum tree problem. For any shown in the figure, 104.5 and 105 , is smaller than 1%. All
fixed k, averages obtained by the pool method should error bars are smaller than the marker sizes.
converge slowly to the true values as the pool size Npool
tends to infinity. Therefore an important question is
whether the pool size is large enough. We focus on the case of true measurements, commenting
To check whether the pool size in our study (106 ) is on the differences for the simpler case of forced measure-
large enough, we choose θ close to the critical point and ments. The forced measurement case falls into the class
study the curves of ln Zktyp with different pool sizes for of tree tensor networks for which order parameter scal-
both the forced and real measurement cases. The result ing was discussed in Ref. [22] using an analogy with an
is shown in Fig. 13. The relative error between the two FKPP-like equation.
curves for the largest pool sizes shown in the figure, 104.5 In the main text (Eq. 55) we discussed the recursion
and 105 , is smaller than 1%. This convinces us that the relation for the generating function Gk that follows from
pool size of 106 used in the main text is sufficient. the linearized recursion relation for Z. We now give the
nonlinear form of the recursion that allows for (1) higher-
order terms in Zk,1 , Zk,2 in Eq. 51 and (2) a dependence
Appendix E: Order parameter scaling of the outcome probability p(s) on Zk,1 and Zk,2 .
In this Appendix (only), we will use U to denote the
In this Appendix we discuss how to use the recursion collection of random unitaries involved in a given node.
relation for Gk to extract the scaling of the order param- Below, various quantities (Ai , F , p) depend both on U
eter in the entangling phase (but close to the transition). and on s, but in order to shorten the formulas we will
24

often suppress these arguments. However we will make When θ = θc , this equation has the exact solution (for
explicit all dependency on the quantities Z1 ≡ Zk,1 and all x) Hk+1 (x) = Hk (x) = exp(−x).15
Z2 ≡ Zk,2 . Let us consider the entangled phase, where
We begin with the nonlinear version of Eq. 51: G ≡ limk→∞ Gk has a nontrivial limit, satisfying
Eq. E5 with Gk = Gk+1 = G. If we are very close to the
Zk+1 = Ai Zi + F (Z1 , Z2 ), (E1) phase transition then Z typ is very small, and there is an
“intermediate regime” of negative x where
where the expansion of F starts at order Z 2 . The sum
over i = 1, 2 is assumed in the first term. Ai and F de- 1  |x|  | ln Z typ |. (E7)
pend on the set s of measurement outcomes at the node,
and the probability distribution ps for s also depends on Then, in addition to linearizing the equation in H (as
Z. To emphasize the Z dependence of ps we will write it above), we may also treat ex as a small quantity. Ne-
as: glecting all terms of order ex in (E6) gives16
* +
ps (Z1 , Z2 ). (E2) X
H(x) = 2 ps (0, 0)H(x − ln A1 ) . (E8)
s
The generating function of interest is
* + This equation was essentially solved in Sec. IV A. We
X
−e−x (Ai Zi +F (Z1 ,Z2 )) use the ansatz H(x) ∝ exp(−λx). Because here we are
Gk+1 (x) = ps (Z1 , Z2 )e .
considering a solution that is stationary (in k), the wave
s U,Z
(E3) parameter λ must be chosen so that the velocity vθ (λ)
In order to express the right-hand side in terms of Gk , vanishes, as discussed in Sec. IV B. This condition gives
we write it as the solution discussed in the main text, with

* + c θc − θ
X
−e−x F (Z1 ,Z2 ) − i e−xi Ai Zi
P
1 + ∂x ln H = √ . (E9)
ps (Z1 , Z2 )e e . tan(φ + c θc − θ x)


s

U,Z xi →x
As discussed in the main text, a heuristic matching argu-
We see that in the first two factors we can make the ment at the left-hand-side of the range of x under√discus-
replacement sion allows ln Z typ be expressed in terms of φ/c θc − θ.
The key point is that the second term in (E9) should
1 ∂ exi ∂ become non-negligible (and positive) as x approaches
Zi → − = . (E4)
Ai ∂e−xi Ai ∂xi ln Z typ , and this matching requires the argument of the
tangent to vanish close to the front.
Then we can average over the Zi , giving Next we consider the value of φ. In the case of forced
X measurements the value of φ can be fixed straightfor-
ps A−1 x1 −1 x2

Gk+1 (x) ' 1 e ∂1 , A2 e ∂2 wardly (if non-rigorously) by matching on the other side
s of the intermediate-x range. For forced measurements, a
× exp −e−x F (A−1 x1 −1 x2 formula like Eq. E9 holds, but with the first term being

1 e ∂1 , A2 e ∂2 )
 −1/2 instead of −1. Therefore the slope in the interme-
diate regime is close to −1/2. On the other hand, it is

× Gk (x1 − ln A1 )Gk (x2 − ln A2 ) .
U xi →x easy to show (for example from the definition of the gen-
(E5) erating function) that for x  1 the slope must approach
−1. Therefore, for forced measurements, we argue that
This equation is in principle exact. For the case of forced φ = π, in order that, as x approaches 0 from the left,
measurements we would have a similar equation, without the slope
the sum over s or the probability factor ps . √ should decrease (by an amount much larger
than θc − θ) to match onto the appropriate solution
When x is far to the right of the front for x & 0.
(x − ln Z typ  1), the quantity Hk = 1 − Gk is ex- However for real measurements we have ∂ ln H ' −1
ponentially small in x − ln Z typ , and we can linearize both in the intermediate regime and for x  1. Therefore
the equation in H. Exploiting the symmetry between we cannot use the same argument to fix φ.
Z1 and Z2 , this linearization gives
X
1
ps A−1 x1

Hk+1 (x) ' 1 e ∂1 , 0
2 s
15 One can check this by using Eq. 63, together with the fact that
the expansion of F in its argument starts at quadratic order, and
× exp −e−x F (A−1 x1

1 e ∂1 , 0) ps (Z, Z 0 ) is normalised for all values of Z, Z 0 (which
the fact thatP
 means that s ∂Z ps (Z, 0)|Z=0 = 0).
Note that −e−x F (A−1 −1 x2
16 x1 x
× Hk (x1 − ln A1 ) . (E6) 1 e ∂1 , A2 e ∂2 ) is of order e , because
U x1 →x the Taylor expansion of F starts at quadratic order.
25

Instead, it seems plausible that φ can be fixed to π/2 where B is a sum of terms involving higher moments (and
by more careful considerations along the following lines. higher powers of the first moment). Since 2hAi − 1 is of
We have (we consider the limit k → ∞) order (θc − θ), this equation strongly suggests that, close
to the critical point, the higher moments are smaller than
the first moment by a factor of order (θc − θ):
G(x) = hexp(−e−x Z)i. (E10)
hZ n i/hZi = O(θc − θ) for n ≥ 2. (E13)
Differentiating to obtain an expression for the slope We have confirmed this numerically for n = 2. Applied
∂x ln H discussed above, and defining Z̃ = e−x Z, gives to Eq. E11, this scaling implies that, in the limit where
θc − θ tends to zero, but for any fixed x,
h1 − e−Z̃ − Z̃e−Z̃ i 1 + ∂x ln H ' (θc − θ)f (x), (E14)
1 + ∂x ln H = . (E11)
1 − he−Z̃ i
where f (x) is an unknown function of x. (We have also
checked this relation numerically, using the expression in
Both the numerator and denominator may in principle be Eq. E11.)
expanded in moments of Z to all orders. The key point is Next we would like to compare Eq. E14 with Eq. E9. If
that while the expansion of the denominator starts with in Eq. E9 we fix a value of x, and then √take the limit of
the first moment, hZi, the expansion of the numerator small θc − θ, we obtain 1 + ∂x ln H ' c θc − θ/ tan(φ).
starts with the second moment. At first glance, consistency with (E14) requires 1/ tan(φ)
Directly averaging the recursion relation for Zk , in the to vanish, allowing us to fix φ = π/2. However, to justify
stationary regime where the moments are independent of this conclusion we would need to quantify the size of
k, gives corrections to (E9) arising from the terms (involving ex )
that we neglected in Eq. E6, in order to check that the
two approximations (E9) and (E14) have a nonvanishing
hZi = 2hAihZi + B, (E12) region of overlap at small θc − θ.

[1] Brian Skinner, Jonathan Ruhman, and Adam Nahum, tum nonlocality in critical hybrid circuits,” Phys. Rev. B
“Measurement-induced phase transitions in the dynamics 104, 104305 (2021).
of entanglement,” Phys. Rev. X 9, 031009 (2019). [11] Aidan Zabalo, Michael J. Gullans, Justin H. Wilson,
[2] Yaodong Li, Xiao Chen, and Matthew P. A. Fisher, Sarang Gopalakrishnan, David A. Huse, and J. H.
“Quantum zeno effect and the many-body entanglement Pixley, “Critical properties of the measurement-induced
transition,” Phys. Rev. B 98, 205136 (2018). transition in random quantum circuits,” Phys. Rev. B
[3] Amos Chan, Rahul M. Nandkishore, Michael Pretko, 101, 060301 (2020).
and Graeme Smith, “Unitary-projective entanglement [12] Michael J. Gullans and David A. Huse, “Scalable probes
dynamics,” Phys. Rev. B 99, 224307 (2019). of measurement-induced criticality,” Phys. Rev. Lett.
[4] Yaodong Li, Xiao Chen, and Matthew P. A. Fisher, 125, 070606 (2020).
“Measurement-driven entanglement transition in hybrid [13] Qicheng Tang and W. Zhu, “Measurement-induced phase
quantum circuits,” Phys. Rev. B 100, 134306 (2019). transition: A case study in the nonintegrable model
[5] M. Szyniszewski, A. Romito, and H. Schomerus, “En- by density-matrix renormalization group calculations,”
tanglement transition from variable-strength weak mea- Phys. Rev. Research 2, 013022 (2020).
surements,” Phys. Rev. B 100, 064204 (2019). [14] Yohei Fuji and Yuto Ashida, “Measurement-induced
[6] Soonwon Choi, Yimu Bao, Xiao-Liang Qi, and Ehud Alt- quantum criticality under continuous monitoring,” Phys.
man, “Quantum error correction in scrambling dynamics Rev. B 102, 054302 (2020).
and measurement-induced phase transition,” Phys. Rev. [15] Oliver Lunt and Arijeet Pal, “Measurement-induced en-
Lett. 125, 030505 (2020). tanglement transitions in many-body localized systems,”
[7] Michael J. Gullans and David A. Huse, “Dynamical pu- Phys. Rev. Research 2, 043072 (2020).
rification phase transition induced by quantum measure- [16] M. Szyniszewski, A. Romito, and H. Schomerus, “Uni-
ments,” Phys. Rev. X 10, 041020 (2020). versality of entanglement transitions from stroboscopic to
[8] Yimu Bao, Soonwon Choi, and Ehud Altman, “Theory continuous measurements,” Phys. Rev. Lett. 125, 210602
of the phase transition in random unitary circuits with (2020).
measurements,” Phys. Rev. B 101, 104301 (2020). [17] Xhek Turkeshi, Rosario Fazio, and Marcello Dalmonte,
[9] Chao-Ming Jian, Yi-Zhuang You, Romain Vasseur, and “Measurement-induced criticality in (2 + 1)-dimensional
Andreas W. W. Ludwig, “Measurement-induced critical- hybrid quantum circuits,” Phys. Rev. B 102, 014315
ity in random quantum circuits,” Phys. Rev. B 101, (2020).
104302 (2020). [18] Mathias Van Regemortel, Ze-Pei Cian, Alireza Seif, Hos-
[10] Yaodong Li, Xiao Chen, Andreas W. W. Ludwig, and sein Dehghani, and Mohammad Hafezi, “Entanglement
Matthew P. A. Fisher, “Conformal invariance and quan- entropy scaling transition under competing monitoring
26

protocols,” Phys. Rev. Lett. 126, 123604 (2021). [37] V. Murg, F. Verstraete, Ö. Legeza, and R. M. Noack,
[19] Ruihua Fan, Sagar Vijay, Ashvin Vishwanath, and Yi- “Simulating strongly correlated quantum systems with
Zhuang You, “Self-organized error correction in random tree tensor networks,” Phys. Rev. B 82, 205105 (2010).
unitary circuits with measurement,” Phys. Rev. B 103, [38] P. Silvi, V. Giovannetti, S. Montangero, M. Rizzi, J. I.
174309 (2021). Cirac, and R. Fazio, “Homogeneous binary trees as
[20] Yaodong Li and Matthew P. A. Fisher, “Statistical me- ground states of quantum critical hamiltonians,” Phys.
chanics of quantum error correcting codes,” Phys. Rev. Rev. A 81, 062335 (2010).
B 103, 104306 (2021). [39] Wei Li, Jan von Delft, and Tao Xiang, “Efficient simu-
[21] Oles Shtanko, Yaroslav A Kharkov, Luis Pedro Garcı́a- lation of infinite tree tensor network states on the bethe
Pintos, and Alexey V Gorshkov, “Classical models lattice,” Phys. Rev. B 86, 195137 (2012).
of entanglement in monitored random circuits,” arXiv [40] V. Murg, F. Verstraete, R. Schneider, P. R. Nagy, and Ö.
preprint arXiv:2004.06736 (2020). Legeza, “Tree tensor network state with variable tensor
[22] Adam Nahum, Sthitadhi Roy, Brian Skinner, and order: An efficient multireference method for strongly
Jonathan Ruhman, “Measurement and entanglement correlated systems,” Journal of Chemical Theory and
phase transitions in all-to-all quantum circuits, on quan- Computation 11, 1027–1036 (2015), pMID: 25844072,
tum trees, and in landau-ginsburg theory,” PRX Quan- https://doi.org/10.1021/ct501187j.
tum 2, 010352 (2021). [41] G. Vidal, “Entanglement renormalization,” Phys. Rev.
[23] Sagar Vijay, “Measurement-driven phase transition Lett. 99, 220405 (2007).
within a volume-law entangled phase,” arXiv preprint [42] Brian Swingle, “Entanglement renormalization and
arXiv:2005.03052 (2020). holography,” Phys. Rev. D 86, 065007 (2012).
[24] Lukasz Fidkowski, Jeongwan Haah, and Matthew B [43] Naoki Nakatani and Garnet Kin-Lic Chan, “Efficient
Hastings, “How dynamical quantum memories forget,” tree tensor network states (ttns) for quantum chemistry:
Quantum 5, 382 (2021). Generalizations of the density matrix renormalization
[25] Andrew C. Potter and Romain Vasseur, “Entanglement group algorithm,” The Journal of Chemical Physics 138,
dynamics in hybrid quantum circuits,” (2021). 134113 (2013), https://doi.org/10.1063/1.4798639.
[26] Matthew P. A. Fisher, Vedika Khemani, Adam Nahum, [44] Crystal Noel, Pradeep Niroula, Daiwei Zhu, Andrew
and Sagar Vijay, “Random quantum circuits,” (2022), Risinger, Laird Egan, Debopriyo Biswas, Marko Cetina,
10.48550/ARXIV.2207.14280. Alexey V. Gorshkov, Michael J. Gullans, David A. Huse,
[27] Romain Vasseur, Andrew C. Potter, Yi-Zhuang You, and and Christopher Monroe, “Measurement-induced quan-
Andreas W. W. Ludwig, “Entanglement transitions from tum phases realized in a trapped-ion quantum com-
holographic random tensor networks,” Phys. Rev. B 100, puter,” Nature Physics 18, 760–764 (2022).
134203 (2019). [45] Matteo Ippoliti and Vedika Khemani, “Postselection-free
[28] Patrick Hayden, Sepehr Nezami, Xiao-Liang Qi, entanglement dynamics via spacetime duality,” Phys.
Nathaniel Thomas, Michael Walter, and Zhao Yang, Rev. Lett. 126, 060501 (2021).
“Holographic duality from random tensor networks,” [46] Jin Ming Koh, Shi-Ning Sun, Mario Motta, and
Journal of High Energy Physics 2016, 9 (2016). Austin J. Minnich, “Experimental realization of a
[29] Tianci Zhou and Adam Nahum, “Emergent statistical measurement-induced entanglement phase transition on
mechanics of entanglement in random unitary circuits,” a superconducting quantum processor,” (2022).
Phys. Rev. B 99, 174205 (2019). [47] Yaodong Li, Yijian Zou, Paolo Glorioso, Ehud Altman,
[30] Adam Nahum, Sagar Vijay, and Jeongwan Haah, “Op- and Matthew P. A. Fisher, “Cross entropy benchmark
erator spreading in random unitary circuits,” Phys. Rev. for measurement-induced phase transitions,” (2022).
X 8, 021014 (2018). [48] Samuel J Garratt, Zack Weinstein, and Ehud Altman,
[31] Yaodong Li, Romain Vasseur, Matthew Fisher, and An- “Measurements conspire nonlocally to restructure crit-
dreas WW Ludwig, “Statistical mechanics model for clif- ical quantum states,” arXiv preprint arXiv:2207.09476
ford random tensor networks and monitored quantum cir- (2022).
cuits,” arXiv preprint arXiv:2110.02988 (2021). [49] Aashish A Clerk, Michel H Devoret, Steven M Girvin,
[32] Fergus Barratt, Utkarsh Agrawal, Sarang Gopalakr- Florian Marquardt, and Robert J Schoelkopf, “Introduc-
ishnan, David A Huse, Romain Vasseur, and An- tion to quantum noise, measurement, and amplification,”
drew C Potter, “Field theory of charge sharpening in Reviews of Modern Physics 82, 1155 (2010).
symmetric monitored quantum circuits,” arXiv preprint [50] B. Derrida and H. Spohn, “Polymers on disordered trees,
arXiv:2111.09336 (2021). spin glasses, and traveling waves,” Journal of Statistical
[33] Javier Lopez-Piqueres, Brayden Ware, and Romain Physics 51, 817–840 (1988).
Vasseur, “Mean-field entanglement transitions in random [51] Ronald Aylmer Fisher, “The wave of advance of advan-
tree tensor networks,” Phys. Rev. B 102, 064202 (2020). tageous genes,” Annals of eugenics 7, 355–369 (1937).
[34] Brayden Ware, Romain Vasseur, and Adam Nahum, “In [52] Jeffrey D. Miller and Bernard Derridda, “Weak-disorder
preparation,” . expansion for the anderson model on a tree,” Journal of
[35] Y.-Y. Shi, L.-M. Duan, and G. Vidal, “Classical simula- Statistical Physics 75, 357–388 (1994).
tion of quantum many-body systems with a tree tensor [53] Cécile Monthus and Thomas Garel, “Anderson transition
network,” Phys. Rev. A 74, 022320 (2006). on the cayley tree as a traveling wave critical point for
[36] L. Tagliacozzo, G. Evenbly, and G. Vidal, “Simulation various probability distributions,” Journal of Physics A:
of two-dimensional quantum systems using a tree tensor Mathematical and Theoretical 42, 075002 (2009).
network that exploits the entropic area law,” Phys. Rev. [54] I. Garcı́a-Mata, O. Giraud, B. Georgeot, J. Martin,
B 80, 235127 (2009). R. Dubertrand, and G. Lemarié, “Scaling theory of the
27

anderson transition in random graphs: Ergodicity and [57] Marc Mézard and Andrea Montanari, “Reconstruction
universality,” Phys. Rev. Lett. 118, 166801 (2017). on trees and spin glass transition,” Journal of statistical
[55] Eric Brunet and Bernard Derrida, “Shift in the velocity physics 124, 1317–1350 (2006).
of a front due to a cutoff,” Physical Review E 56, 2597 [58] M. Kieburg, A. B. J. Kuijlaars, and D. Stivigny, “Sin-
(1997). gular value statistics of matrix products with truncated
[56] Antoine Tilloy, Michel Bauer, and Denis Bernard, unitary matrices,” International Mathematics Research
“Spikes in quantum trajectories,” Physical Review A 92, Notices 2016, 3392–3424 (2015).
052111 (2015).

You might also like