<<

Measurement-induced steering of quantum systems

Sthitadhi Roy,1, 2, ∗ J. T. Chalker,1, † I. V. Gornyi,3, 4, ‡ and Yuval Gefen5, § 1Rudolf Peierls Centre for Theoretical , Clarendon Laboratory, Oxford University, Parks Road, Oxford OX1 3PU, United Kingdom 2Physical and Theoretical Chemistry, Oxford University, South Parks Road, Oxford OX1 3QZ, United Kingdom 3Institute for Quantum Materials and Technologies, Karlsruhe Institute of Technology, 76021 Karlsruhe, Germany 4A. F. Ioffe Physico-Technical Institute of the Russian Academy of Sciences, 194021 St. Petersburg, Russia 5Department of Condensed Matter Physics, Weizmann Institute of Science, 7610001 Rehovot, Israel (Dated: June 24, 2020) We set out a general protocol for steering the state of a quantum system from an arbitrary initial state towards a chosen target state by coupling it to auxiliary quantum degrees of freedom. The protocol requires multiple repetitions of an elementary step: during each step the system evolves for a fixed time while coupled to auxiliary degrees of freedom (which we term ‘detector ’) that have been prepared in a specified initial state. The detectors are discarded at the end of the step, or equivalently, their state is determined by a projective measurement with an unbiased average over all outcomes. The steering harnesses back-action of the detector qubits on the system, arising from entanglement generated during the coupled evolution. We establish principles for the design of the system-detector coupling that ensure steering of a desired form. We illustrate our general ideas using both few-body examples (including a pair of spins-1/2 steered to the singlet state) and a many-body example (a -1 chain steered to the Affleck-Kennedy-Lieb-Tasaki state). We study the continuous time limit in our approach and discuss similarities to (and differences from) drive-and-dissipation protocols for quantum state engineering. Our protocols are amenable to implementations using present-day technology. Obvious extensions of our analysis include engineering of other many-body phases in one and higher spatial dimensions, adiabatic manipulations of the target states, and the incorporation of active error correction steps.

CONTENTS V. Discussion 13 A. Experimental realisations 14 I. Introduction1 B. Connection to previous work 15 Structure of the paper3 C. Future directions 16

II. Overview3 Acknowledgments 17 A. Steering protocol3 B. Steering inequalities4 A. Steering a single spin-1/2 to an arbitrary state 17 C. Guiding principle for choice of system-detector coupling Hamiltonian4 B. Derivation of steering inequality with multiple D. Toy example with a spin-1/24 detector qubits 17 E. Many-body states and multiple detectors5 C. Shake-and-steer protocol 18 F. Steady states and rate of approach5 D. Monotonicity of steering to the AKLT state 19 III. Formalism of quantum state steering6 A. General derivation of steering inequality from References 20 discrete-time map6 B. Effective Lindblad dynamics7 C. Steering a pair of spins-1/28 I. INTRODUCTION IV. Spin-1 chain9 arXiv:1912.04292v2 [cond-mat.stat-mech] 23 Jun 2020 A. A pair of spins-19 There are many circumstances in which one would like B. Steering to the AKLT state 10 to initialise a many-body quantum system in a specified state: examples of current interest range from processing to studies of non-equilibrium dy- namics. Two standard approaches for preparing quan- ∗ [email protected] tum states are suggested by the laws of quantum me- † [email protected] chanics and statistical mechanics. One of these is to ‡ [email protected] make projective measurements of a set of observables § [email protected] represented by commuting operators that fully specify 2 the target state. Alternatively, if the target state is the steering has been narrowed to the impossibility of local ground state of the Hamiltonian for the system, it can be hidden state models describing the ensemble of states reached by putting the system in thermal contact with a that a system can take upon measurement of another heat bath that is at a sufficiently low temperature. Both system entangled with it [4–6]. We however continue to approaches have disadvantages, especially for a system use the term in its broader sense. with a large number of degrees of freedom: in the first To set out in more detail the links between our protocol approach, a general initial state is not an eigenstate of for preparing or steering a quantum state and discussions the measurement operator and hence the measurement of quantum measurement, it is useful to recall the distinc- outcome is probabilistic. The probability that the tar- tion between strong or projective measurements and the get state is reached decreases rapidly towards zero with notion of weak quantum measurements [7–13]. The for- increasing system size. With the second approach, the mer destroy the coherent quantum dynamics of the sys- temperature scale required to completely eliminate ther- tem. By contrast, under the latter, auxiliary degrees of mal excitations from a system also decreases towards zero freedom are coupled weakly to the system and projective with size. In principle, if the initial state of the system is measurements are performed on these detector qubits af- precisely known, a further possibility is to act on the state ter decoupling them from the system. No matter how with an appropriately chosen perturbation for a precise weak the coupling, a measurement of this kind unavoid- interval of time, so that it evolves into the target state; ably impacts the system through its back-action, even this, however, requires extreme fine-tuning in a large sys- after the detector qubits are decoupled from the system. tem. Instead of viewing the measurement-induced disturbance Our aim in the following is to establish a class of pro- on the system’s state as a handicap, it may be consid- tocols for quantum state preparation that improve on ered a resource for quantum control and manipulation. both projective measurement and thermal contact with The back-action of measuring the state of detectors en- a heat bath. A protocol of the type we describe will, tangled with the system can be harnessed to control the in an ideal implementation, steer a system from an ar- system’s evolution and steer its state towards a desired bitrary initial state to the target state, with guaranteed target state: this constitutes the central message of our success. It does so by coupling the quantum system of work. In principle, one could make use of the measure- interest to external detector qubits (auxiliary quantum ment outcomes to determine whether the state of the sys- degrees of freedom) that have been prepared in specified tem has been successfully steered in a realisation of the initial states, then evolving the coupled system under experiment, and discard it if this is not the case. In the standard unitary quantum dynamics for a fixed time in- simplest version of such postselected protocols, the prob- terval, and finally decoupling and discarding the detector ability of success then goes down exponentially with the qubits. Multiple repetitions of this process using freshly length of the outcome sequence or the number of detec- prepared detector qubits on each occasion, coupled to the tors. This suppression is also likely to be exponentially system with a suitable interaction Hamiltonian, produce strong in system size for a many-body system. To avoid the outcome we require. In this paper we illustrate our such suppression, our aim is to establish protocols within general approach both for small systems consisting of one a setting where the outcomes are averaged over in an un- or two spin-1/2 degrees of freedom and for a macroscopic biased fashion, and hence there is no loss of probability. many-body system. We refer to processes of this type as blind measurements This protocol and possible generalisations have the or non-selective measurements. fundamental concepts of and The concept of weak measurements allows for the no- quantum measurements as essential ingredients, and pos- tion of continuous measurements [14–16]. The latter can sess conceptual links to the theory of open quantum sys- be viewed as a discrete sequence of weak measurements tems. First, the effect of a coupling between the sys- in the limit of vanishing time-interval for each weak mea- tem and detector qubits is to generate entanglement be- surement. This results in a sequence of measurement out- tween their states. Second, focussing on the behaviour comes defining a quantum trajectory. The equation for of the system after discarding the detector qubits, this the density matrix of the system averaged over the tra- entanglement induces steering of its quantum state, of jectories is governed by a Lindblad equation. In our case a kind first discussed by Schr¨odingerin the context of the steering protocol consists of a discrete sequence of the Einstein-Podolsky-Rosen (EPR) paradox [1,2]. A measurements on detector qubits not necessarily coupled physical setting for the decoupling and discarding of the weakly to the system. A suitably defined time-continuum detectors is that of performing strong measurements on limit maps the dynamics of the density matrix of the sys- these qubits, and then taking an unbiased average over tem to a Lindblad equation. For part of the analysis in the outcomes. Indeed, Schr¨odinger’soriginal formula- our work, we exploit this formal connection between our tion of steering [3] was in terms of a sophisticated exper- steering protocol and Lindblad dynamics to gain insight imenter performing suitable measurements on one of the into the steady state and the rate of approach to it. parts of a bi-partite system to drive the other part to a The emergence of Lindblad dynamics suggests a com- state, chosen by the experimenter, with non-zero prob- parison of our protocol with proposals for preparing ability. More recently, the scope of the term quantum non-trivial many-body states in open quantum systems, 3 where the dissipative environment is posited to be Marko- Structure of the paper vian and is treated within the framework of Lindblad dynamics [17–26]. The challenge in that context is to The rest of the paper is structured as follows. We start find suitable Lindblad jump operators that can be cast with an overview in Sec.II, where we introduce the basic in terms of physical systems participating in the dynami- ingredients of the steering protocol and state the main cal process. Our measurement-induced steering protocol results of the work. SectionIII presents the details of has some technical similarities with this framework, as the steering protocol. The guiding principle for design- well as conceptual distinctions from it. With regard to ing the protocol is introduced and derived in Sec.IIIA, the former, if the detector qubits we discuss are viewed as followed by the derivation of the effective Lindblad dy- an environment, this environment is Markovian by con- namics in Sec.IIIB. The workings and advantage of the struction, since the detector qubits are prepared afresh approach is exemplified via the simple case of a spin- at the start of every cycle. Hence, the map govern- 1/2 pair steered towards their singlet state in Sec.IIIC. ing the evolution of the density matrix of the system We then turn to a many-body quantum system, a spin- has a representation in terms of Kraus operators, which 1 chain – which we steer to the Affleck-Kennedy-Lieb- ultimately leads to a Lindblad equation in the time- Tasaki (AKLT) ground state [33, 34] in Sec.IV. As a continuum limit [27, 28]. building block of the protocol, the steering of a spin-1 An obvious advantage of our protocol is that the pair is discussed in Sec.IVA, followed by the numerical jump operators in the emergent Lindblad equation are treatment of the spin-1 chain in Sec.IVB. We close with uniquely and automatically fixed by the details of the discussions of possible experimental implementations of system-detector coupling Hamiltonian, thus facilitating the protocol and future directions in Sec.V. “on-demand” engineering of Lindbladians. On the other hand, since our protocol is fundamentally a microscopic measurement protocol, certain requirements can easily II. OVERVIEW be relaxed to go far beyond what is describable within the standard Lindblad framework. We note earlier works in the context of quantum control, discussing manipula- The central result of this work is a general formalism tion of quantum states via repeated interactions with de- for measurement-induced steering of a many-body quan- signed environments [23, 29, 30]; much of this, however, tum system towards a non-trivial target state, by repeat- has focussed on the formal aspects of the theory or on edly coupling and decoupling a set of auxiliary degrees of applications restricted to few-body systems. freedom interspersed with unitary dynamics of the com- Another advantage of a measurement-based protocol posite system. The protocol makes use of the entangle- over coupling with an environment is that, whereas the ment generated between the system and the auxiliary latter simply gets entangled with the system, a detector degrees of freedom to steer the state. The auxiliary de- can be read out, yielding information on the system grees of freedom are simple quantum systems with small state which could be further used to accelerate conver- Hilbert-space dimensions such that they are easy to pre- gence to the desired state. Obvious examples include the pare in a desired initial state; in this work we consider use of post-selected protocols or using the measurement a set of decoupled detector qubits (spins-1/2) initially outcomes to implement active error correction and en- polarised along a given direction. hance the blind measurement protocols [31, 32]. To summarise, our measurement-based steering pro- tocol differs fundamentally from and has various advan- A. Steering protocol tages over related protocols for manipulating and sta- bilising non-trivial quantum states. First, the measure- The steering protocol can be described as a sequence ments in our case need not be weak: the detectors may be of discrete steering events each of which consists of the strongly coupled to the system. Second, in an appropri- following steps: ate time-continuum limit the dynamics is described by a Lindblad equation, but rather than introducing a Marko- The detector qubits are prepared in a initial given vian reservoir or jump operators “by hand”, bath cou- state, which does not depend on the state of the pling is engineered by the measurement operators, which system. We denote the initial state of the detector fix the jump operators uniquely. Third, unlike unstruc- qubits as Φd and the density matrix correspond- | i tured environments or baths, the detector qubits can be ing to this initial state as ρd. read out, with the outcomes exploited to further the ef- ficiency and scope of our steering on platforms employ- The system is then coupled to the detector qubits ing postselected protocols or using active decision mak- and the composite system evolves unitarily under a ing protocols based on the read-out sequence. Finally, Hamiltonian for some time. Denoting the density we note the non-triviality of applying our measurement- matrix of the system at time t as ρs(t) and the based steering to many-body systems using an extensive system-detector Hamiltonian as Hs-d, the state of number of auxiliary degrees of freedom. the system-detector composite after evolution for 4

an interval δt is given by see from this that for the inequality to be as strong as possible, W should be the swap operator between these −iHs-dδt iHs-dδt ρs-d(t + δt) = e ρd ρs(t)e . (1) two spaces and ρ should be the image of ρ under this ⊗ d s⊕ swap. More generally, a Hamiltonian which contains di- The detector qubits are then decoupled from the rect products of operators O(n) in the detectors’ subspace system. Formally we may trace out the detector d d that rotate the detectors from their initial state to an degrees of freedom to obtain the density matrix of H (n) orthogonal subspace and operators U in the system’s the system at time t + δt s subspace s that rotate the system to the target state H ρs(t + δt) = Trdρs-d(t + δt). (2) manifold from an orthogonal subspace, will steer the sys- tem towards the target state. Such a Hamiltonian has Equations (1) and (2) describe the discrete time- the form

evolution map for the density matrix of the system X  (n)  (n) under the steering dynamics. Hs-d = Od Φd Φd Us + h.c., (7) n | i h | ⊗

The detector qubits are re-prepared in their initial (n) states and the above steps are repeated. where n labels the detector qubit. Since Od connects the state Φd to its orthogonal subspace, it satisfies (n) | i Φd Od Φd = 0. Additionally, we assert the following B. Steering inequalities hproperties| | fori the system operators: (n) (n) (i) Us annihilates the target state, Us Ψs = 0 As we would like the system to get steered towards the | ⊕ i target state, denoted by Ψs⊕ (with corresponding den- (n) (n) (n)† | i (ii) Us is normalised such that Us Us Ψs⊕ = sity matrix ρs ), the dynamics induced by Hs-d should | i ⊕ Ψs satisfy the steering inequality | ⊕ i (n) (m) (iii) for different n and m, if Us and Us share spatial Ψs⊕ ρs(t + δt) Ψs⊕ Ψs⊕ ρs(t) Ψs⊕ , t, (3) (m) (n)† h | | i ≥ h | | i ∀ support then Us Us = 0. If they do not share with the inequality ideally becoming an equality only if a spatial support, they automatically commute. ρ (t) = ρ . s s⊕ To ensure that equality in Eq. (3) always implies ρ (t) = Note that this inequality is very strong. When com- s ρs⊕ it is necessary that Ψs⊕ is coupled to every state bined with the condition on equality, they ensure that | i (n)† the system is steered to the target state irrespective of in its complement by at least one operator Us . In our its initial state. In principle, one could envisage weaker study of steering to the AKLT state, we observe guar- steering inequalities. One example is anteed convergence to the target state despite the set of (n)† operators Us not connecting the AKLT state to all lim Ψs ρs(t) Ψs = 1 (4) { } t→∞ h ⊕ | | ⊕ i the other states in the Hilbert space. However, mono- tonicity is not ensured and, in principle, can depend on and a still weaker one is the particular measure of distance from the target state. We disucss this issue in detail in Secs.IIIA,IVB and lim Ψs ρs(t) Ψs > Ψs ρs(0) Ψs . (5) t→∞ h ⊕ | | ⊕ i h ⊕ | | ⊕ i AppendixD. We present in Sec.IIIA a general strategy for designing system-detector couplings so that the strongest of these D. Toy example with a spin-1/2 forms, (3), holds. As a simple example consider a single spin-1/2 which C. Guiding principle for choice of system-detector we wish to steer to the fully polarised state along the coupling Hamiltonian z-direction, so that Ψs⊕ = s . The subspace orthog- | i |↑ i † onal to Ψs has only a single state: U Ψs = s . | ⊕ i s | ⊕ i |↓ i With the protocol fixed, it remains to find a guiding Hence, a single detector qubit is sufficient to steer the principle for the choice of the Hamiltonian that governs state. Without loss of generality, let us choose it to be initialised as Φd = d . Naturally, we also have the evolution of the system-detectors composite. As a | i |↑ i Od Φd = d . Following Eq. (7), the system-detector first step, consider summing both side of (3) over ρs(t) | i |↓ i representing all possible pure states. If (3) is obeyed, coupling Hamiltonian can then be written as −iHs-dδt then W e must satisfy Hs-d = J( d d ) ( s s ) + h.c. ≡ |↓ i h↑ | ⊗ |↑ i h↓ | (8) † − + + − Trs,d[W ρd 11sW 11d ρs ] Trs,d[ρd ρs ] . (6) = J σd σs + σd σs , ⊗ · ⊗ ⊕ ≥ ⊗ ⊕ ± In the special case where the Hilbert spaces of the system where σs(d) denote the Pauli raising and lowering opera- ( s) and the detector ( d) have the same dimension, we tors for the system and detector qubits. H H 5

At time t, the state of the system can be generally The particular example we consider in detail is the written as ρs(t) = (I2 +s(t) σs)/2 where s(t) is a classical Affleck-Kennedy-Lieb-Tasaki (AKLT) state of a one- three-component vector. The· state of the system spin- dimensional spin-1 chain [33, 34]. The AKLT state is a 1/2 at time t + δt can then be obtained by evolving the valence bond state such that on each bond between two z combined state of the system and the detector, (I2 +σd) neighbouring spins-1, there is no projection on the total ⊗ ρs(t)/2, with the unitary operator W , and tracing over spin-2 sector. Thus one can steer the spin-1 chain by the detector. This yields the relation for s(t + δt) locally steering each bond out of the total spin-2 sector, and the uniqueness of the ground state of the AKLT chain t sx(t + δt) = cos(Jδt)sx(t) = cos δt (Jδt)sx(0), (with periodic boundary conditions) guarantees steering t to the AKLT state. Steering each bond requires only a sy(t + δt) = cos(Jδt)sy(t) = cos δt (Jδt)sy(0), finite number of detectors due to the finite dimension of 2 2 (9) sz(t + δt) = 1 cos (Jδt) + cos (Jδt)sz(t) the total spin-2 subspace for a pair of spins-1 and hence − 2t 2t the total number of detectors needed to steer a chain is = 1 cos δt (Jδt) + cos δt (Jδt)sz(0). − only extensive in system size. Further examples of states which satisfy the above cri- The above equation explicitly shows that teria include matrix product states, projected pair en- tangled states, the Laughlin state of a fractional quan- lim s(t) = (0, 0, 1), (10) t→∞ tum Hall system [36], the ground state of Kitaev’s [37], and of recent interest, fermionic symmetry- irrespective of the initial conditions and that the limit is protected topologically ordered eigenstates of full com- approached exponentially in time. muting projector Hamiltonians [38, 39]. This example illustrates how a system-detector cou- An obvious potential concern arises when our proto- pling Hamiltonian of the form in Eq. (7) leads to the col is used to steer many-body systems to eigenstates of satisfaction of the desired steering inequality, Eq. (3), for local projector Hamiltonians, because in the cases of in- generic initial states of the system. This guarantees suc- terest a given quantum degree of freedom appears in more cessful steering to the target state from arbitrary initial than one projector. In particular, in such a scenario it is (n) (m) states. In turn, the example shows how weak measure- not guaranteed that two system operators Us and Us ments can produce control and manipulation that pro- acting on such a shared degree of freedom, while steering jective measurements cannot [35]. their corresponding parts of the system locally, satisfy the We remark that although in the above example the conditions listed at the end of Sec.IIC. Steering towards target state of the system qubit is the same as the initial the target space of one projector may therefore undo, at state of the detector qubit, there is nothing special about least partially, the effect of steering to the target space of this choice: see AppendixA. a different projector with which the degree of freedom in question is shared. Despite these complications, we show that our approach enables us to select zero-energy eigen- E. Many-body states and multiple detectors states of local projector Hamiltonians as target states and unique stationary states of our dynamics. We illustrate In the context of many-body systems, the subspace or- this numerically for the AKLT chain, showing steering to thogonal to the target is exponentially large in the system the ground state with guaranteed success from arbitrary size. Hence it may appear that the system-detector cou- initial states. pling Hamiltonian of the form in Eq. (7) entails an expo- nentially large number of detectors, or non-local system- detector coupling Hamiltonians, or both. For implemen- F. Steady states and rate of approach tation of the protocol to be feasible, we would like to have at most an extensive number of detectors, with cou- Equations (1) and (2) show that the time-evolution of plings that are local. Although these constraints appear the density matrix of the system is governed by a linear rather restrictive, they can be satisfied for versions of our map. On general grounds, the largest eigenvalue magni- protocol that allow steering to a large class of strongly tude for this map is unity. If the associated eigenvector correlated quantum states which are eigenstates of un- is unique, this ensures that there exists a unique steady frustrated local projector Hamiltonians. One can steer state. The logarithm of the next eigenvalue then encodes to such states by locally steering different parts of the the rate of approach in time to this steady state. Ideally, system to the appropriate eigenstate of the correspond- we would like the rate to be finite in the thermodynamic ing local projector part of the Hamiltonian. Since the limit so that the time-scale required to steer the many- effective Hilbert dimension of a local part of the system body state arbitrarily close to the target state does not is finite, one can steer locally with a finite number of de- diverge with system size. tectors, and hence the total number of detectors required In the context of the AKLT state, we find that the scales linearly with system size, and not exponentially. magnitude of these eigenvalues is most conveniently stud- Moreover, the couplings are manifestly local. ied in the time-continuum limit (δt 0) of the map in → 6

Eq. (2) where an effective Lindblad equation emerges for shown by noting that with Hs-d from Eq. (7), Hs-d (ρd · ⊗ the equation of motion of ρs with the jump operators be- ρs⊕ ) = 0, which naturally implies ing fixed by the system-detectors coupling Hamiltonian −iHs-dδt iHs-dδt Hs-d. Unit eigenvalue for the discrete map corresponds e ρd ρs⊕ e = ρd ρs⊕ . (11) to zero eigenvalue for the Lindbladian operator. If the ⊗ ⊗ corresponding Lindbladian eigenvector is unique, it is the Upon taking the trace over the detectors, one recovers −iHs-dδt iHs-dδt stationary state of the steering protocol. Successful steer- that Trd[e ρd ρs⊕ e ] = ρs⊕ and hence ρs⊕ ⊗ ing implies that the steady state of the Lindbladian is the is a stationary state of the map. target state. Moreover, the gap in the spectrum of the We next show that the system is steered towards the Lindbladian between the zero eigenvalue and the rest of target state at every discrete steering event. For brevity, the eigenvalues determines the rate of the of approach of we show here the derivation with a single detector qubit the system to the target state. In the case of the AKLT and state the results for multiple detector qubits. De- state, analysis of the size-dependence of the numerically tails of the derivation for the latter are presented in Ap- obtained gap in the effective Lindbladian’s spectrum in- pendix.B. dicates that it stays finite in the thermodynamic limit. To proceed, we find it most convenient to partition the composite Hilbert space of the detectors and the system, d s into two subspaces, which we denote as d s H ⊗H D ⊗H and d s where d is the one-dimensional subspace III. FORMALISM OF QUANTUM STATE D ⊗ H D spanned by Φd and d d = d. In this convention, STEERING | i D ⊕ D H the density matrix ρd ρs(t) can be represented as ⊗ Dd Dd In this section, we describe the quantum steering pro- z }| { z }| {  tocol in detail. We show explicitly that a system detector ρs(t) 0 d ρd ρs(t) = D . (12) coupling Hamiltonian of the form given in Eq. (7) leads to ⊗ 0 0 d D dynamics that satisfy the steering inequality, Eq. (3), on rather general grounds. Additionally, we make the con- Additionally, the system-detector coupling Hamiltonian, nection to an effective Lindblad equation that describes Eq. (7) takes the form the time-continuum limit of the steering map, Eq. (2). Dd Dd  z }| { z }| { 0 U † d Hs-d = D . (13) U 0 d A. General derivation of steering inequality from D discrete-time map The unitary time-evolution operator generated by the system-detector coupling Hamiltonian is

Let us discuss in some generality the map governing ∞ 2k k k ( iδt)  U †U −iδt U † UU †  the discrete time evolution of the system density matrix, −iHs-dδt X ( ) ( 2k+1 ) ( ) e = − k k . (2k)! ( −iδt )U(U †U) (UU †) Eq. (2), with the system-detector coupling Hamiltonian k=0 2k+1 Hs-d of Eq. (7). We will first show that ρs = ρs⊕ = (14)

Ψs⊕ Ψs⊕ is indeed a stationary state of the map and Following the steering map of Eq. (2), applying the above | i h | then show that the system is progressively steered to- unitary matrix to the density matrix of Eq. (12) and wards the ρs⊕ after each steering event. taking the trace of the detector, one obtains the time

That ρs⊕ is a stationary state of the map is easily evolved density matrix of the system as

2k 2l 2k+1 2l+1 X ( iδt) (iδt) h † k † li X ( iδt) (iδt) h † k † †li ρs(t + δt) = − U U ρs(t) U U + − U U U ρs(t)U UU . (15) (2k)!(2l)! (2k + 1)!(2l + 1)! k,l k,l

† Using Us Ψs⊕ =0 and UsUs Ψs⊕ = Ψs⊕ , the diagonal responding to the target state can be obtained from matrix element| i of the time-evolved| i density| i matrix cor- Eq. (15) as

† 2 Ψs⊕ ρs(t + δt) Ψs⊕ = Ψs⊕ ρs(t) Ψs⊕ + Ψs⊕ Usρs(t)Us Ψs⊕ sin (δt). (16) h | | i h | | i |h | {z | }i Q 7

Note Q is a diagonal matrix element of a valid density B. Effective Lindblad dynamics matrix ρs(t) which ensures that it is non-negative. The change in the diagonal element of the density matrix cor- We now show that the time-continuum limit of the responding to the target state is proportional to the sup- map in Eq. (2) leads to a Lindblad equation for the dy- port of the density matrix on the subspace orthogonal to namics of the density matrix of the system. The exposi- Ψs . Hence, the inequality in Eq. (3) becomes an equal- | ⊕ i tion of the conceptual connection between measurement- ity when ρs(t) has no remaining support on the subspace induced quantum steering and the dynamics being de- orthogonal to Ψs – in other words, ρs(t) has reached its | ⊕ i scribed by the Lindblad equation, often associated with target form. This concludes the derivation of the steering dissipative dynamics, is an important result of this work. inequality, Eq. (3), which in turn shows that the system The Lindblad equation is a master equation for the den- is steered towards the target state progressively at each sity matrix of the form steering event. In the case of multiple detectors, the equation for the ∂tρs = [ρs] L   time-evolved density matrix has the form as in Eq. (16) X † 1 † = i[Hs, ρs] + LiρsL L Li, ρs , (20) but Q is given by (see AppendixB for derivation) − i − 2{ i } i

X (n) (n)† where Hs is the intrinsic Hamiltonian of the system and Q = Ψs⊕ Us ρs(t)Us Ψs⊕ (17) n h | | i the Lis are the quantum jump operators. They are fixed in our derivation by the form of the system-detector cou- which is again a sum of diagonal elements of a density pling Hamiltonian. matrix and hence is non-negative. Also it is the total sup- As in Sec.IIIA, we sketch the derivation with a single port of the density matrix on the subspace orthogonal to detector and state the result for multiple detectors. To (n)† Ψs spanned by the set of states Us Ψs . Pro- proceed with the derivation we again partition the com- | ⊕ i { | ⊕ i} vided Ψ is connected to every state in its complement posite Hilbert space and represent ρd ρs(t) as in (12). s⊕ ⊗ | i (n)† The general system-detector coupling Hamiltonian can by an operator Us then Q = 0 implies ρs(t) = ρs⊕ . For a many-body system, if a sufficiently large number be represented by a matrix of the form Dd Dd of operators is employed that the target state is con- z }| { z }| { (n)†  †  d nected to all others by the operators Us , monotonic V U Hs-d = 0 D . (21) convergence to the target state is straightforwardly guar- UV d anteed. The required number of operators may, however, D Expanding the map in Eq. (2) to second order in δt one be exponentially large in system size. By contrast, we obtains are interested in steering a many-body system using only ρs(t + δt) ρs(t) 1 an extensive number of operators, each coupled locally − =i[V, ρs(t)] [V, [V, ρs(t)]]δt+ to the system. It can then happen that Q = 0 for some δt − 2   ρs(t) at a given step, but that further steps successfully † 1 † Uρs(t)U U U, ρs(t) δt. steer the system to the target. In Sec.IVB we show − 2{ } how in a such a setting with extensive, local steering op- (22) erators a spin-1 chain is steered uniquely to the AKLT ˜ state. In this case we also examine whether or not steer- Taking the limit δt 0 with U = U√δt while requiring V (1) and U˜→ (1) [40], one obtains a time- ing is monotonic. We show that this may depend on the || || ∼ O || || ∼ O choices of initial state and of the measure of the distance continuum equation of motion   between the time-evolving state and the target state; see † 1 † ∂tρs(t) = i[V, ρs(t)] + Uρ˜ s(t)U˜ U˜ U,˜ ρs(t) . AppendixD for details. We will quantify the distance of − − 2{ } ρs(t) from the target state, ρs⊕ , via the Frobenius and (23) trace norms, denoted as DF and D1 respectively. The Comparing with Eq. (20), it is trivial to see that the are defined as equation of motion for ρs is indeed of the Lindblad form q with the jump operator given by U˜. The general Hamil- 2 DF (t) = Tr[ρs(t) ρs ] , (18a) − ⊕ tonian in Eq. (21) reduces to the specific one in Eq. (7) hq i with the choice of operators V = V 0 = 0. Hence the ef- 2 D1(t) =Tr (ρs(t) ρs ) /2. (18b) − ⊕ fective Lindblad equation which describes the dynamics of the steering protocol in the time-continuum limit con- Additionally, since the AKLT ground state is an unique sists only of the jump operators. In the case of multiple zero-energy ground state of the AKLT Hamiltonian (with detectors, the Lindblad equation obtained has multiple periodic boundary conditions), another good measure for jump operators the distance is the energy of the state measured with   X ˜ ˜ † 1 ˜ † ˜ respect to the AKLT Hamiltonian ∂tρs = i[V, ρs(t)] + Unρs(t)Un UnUn, ρs(t) . − n − 2{ } EAKLT(t) = Tr[HAKLTρs(t)]. (19) (24) 8

A few comments regarding the Lindblad equation are Jt = 0.0 Jt = 0.1 Jt = 1.0 Jt = 10.0 (a1) (a2) (a3) (a4) 1 in order which will eventually prove useful in the anal- S0 ysis of the concrete examples we discuss. Note that the T+ Lindblad equation is linear in ρs. Hence, if the H - T0 D s dimensional density matrix ρs(t) is unravelled as a su- T 2 − 0 pervector, ~ρs(t), of dimension H , the time-evolution is S0 T+ T0 T S0 T+ T0 T S0 T+ T0 T S0 T+ T0 T D s − − − − simply generated by the superoperator corresponding to 1.0 (b) 8 (c) 100 (d)

the Lindbladian ~ 1 0.8 10− 6

S0 ρs(t) S0 ) ~ h | | i 2 ~ρs(t) = exp ( t) ~ρs(0). (25) 0.6 δt 10− T+ ρs(t) T+ ](

L · h | | i v 4 3 T0 ρs(t) T0 10−

The steady state manifold of the dynamics is then given 0.4 h | | i Re[ − ~ T ρs(t) T h −| | −i 2 10 4 D1 by the manifold of states corresponding to zero eigenval- 0.2 −

∆(δt) distance from singlet DF ues of ~. Hence the uniqueness of the steady state can 5 10 ∆t ~ − L e− be determined by studying the degeneracy of the zero 0.0 0 ∼ 1 0 1 10− 10 10 eigenvalue index 0 25 50 eigenvalues. Moreover, the gap in the spectrum of ~ be- Jt Jt tween the zero eigenvalue and the rest of the eigenvalues,L v ,defined as ∆ = minv6=0 Re[v] determines the na- { } {| |} FIG. 1. Steering of a pair of spins-1/2 to the singlet ture of the approach to the steady state. For a finite ∆ state using three detector qubits coupled via the Hamilto- in the thermodynamic limit, the system approaches the nian in Eq. (28). (a) The density matrix of the system, ρs steady state exponentially fast in time with a rate ∆. in the basis spanned by the singlet and the three triplets as In some of the examples we study, the Lindblad equa- a colourmap (for the absolute values of the matrix elements) tion that arises has a special form. Specifically, if there at different times t. It shows the decaying support on the tot exists a choice of basis α such that all the jump op- S = 2 diagonal elements. (b) The evolution of the diagonal {| i} elements of ρs(t) shows the steering to the singlet. (c) The erators are of the form √γαβ α β , then the matrix el- ements of the density matrix| ini hthe| same basis follow a spectrum of the corresponding Lindbladian with a unique zero eigenvalue (steady state) and a finite gap ∆. (d) The distance master equation given by of ρs(t) from the singlet state |S0i hS0| as measured via the X 1 X matrix norms in Eq. (18) decays exponentially in time at a ∂tρs,αβ = δαβ γαν ρs,νν ρs,αβ (γνα +γνβ), (26) rate is correctly given by ∆. For the plots, J = 1 and δt = 0.1. ν − 2 ν which directly implies that the off-diagonal elements of coupling Hamiltonian motivated from the form in Eq. (7) the density matrix decay exponentially in time. More- can be written as over, the equations for the diagonal elements do not in- volve the off-diagonal elements and hence follow a classi- 3 X −  cal master equation. Hs-d = J σ ρd Ui + h.c. , (28) di ⊗ i=1

C. Steering a pair of spins-1/2 with

1 z − − z We next illustrate these ideas using a two-spin prob- U1 = S0 T+ = 3 [(1 + σ1 )σ2 σ1 (1 + σ2 )], (29a) | i h | 2 2 − lem. Consider a pair of spins-1/2, acted on by the set µ µ 1 z + + z of Pauli matrices, σ1 and σ2 , which we would like U2 = S0 T− = 3 [(1 σ1 )σ2 σ1 (1 σ2 )], (29b) { } { } | i h | 2 2 − − − to steer to the singlet state S0 = ( )/√2.  z z  | i |↑↓i − |↓↑i 1 + − − + σ1 + σ2 The Hilbert space of two spins-1/2 is four-dimensional U3 = S0 T0 = σ1 σ2 σ1 σ2 + . (29c) with the subspace orthogonal to S0 spanned by the | i h | 2 − 2 | i three triplet states, T+ = , T− = , and | i |↑↑i | i |↓↓i † √ Note from the above equations that UiUj = 0 for i = j T0 = ( + )/ 2. Hence, the simplest steering 6 |protocoli |↑↓i uses three|↓↑i detector qubits, each for steering the as each of the triplet states are orthogonal to each other state out of one of the triplet states onto the singlet state. and also to the target singlet state, and hence the con- At the start of every steering event, each of the detector ditions asserted for the system operators in Sec.IIC are qubits is prepared in a pure state fully polarised along satisfied. Results for the evolution of the two-spin sys- the positive z-axis such that tem with the above steering protocol are shown in Fig.1. All the elements of density matrix in the basis spanned z z z I2 + σd1 I2 + σd2 I2 + σd3 by the singlet and the three triplet states decay to zero ρd = , (27) 2 ⊗ 2 ⊗ 2 exponentially in time except for the diagonal element cor- responding to the singlet, which approaches unity. This µ where σdi denotes the set of Pauli matrices for the shows that the system is indeed steered to the singlet th { } i detector. For this choice of ρd, the system-detector state exponentially in time. 9

Turning to the effective Lindblad dynamics corre- The total spin on a bond between the sites ` and ` + 1 tot sponding to the protocol, the three jump operators of will be denoted by S(`,`+1) = S` + S`+1. In what fol-

the Lindblad equation can be read off from Eq. (28) as lows, it will be often convenient to use the simultane- ~ 2 L = JU √δt. The spectrum of the so-obtained Lind-    tot,z  i i ous eigenstates of Sˆtot and Sˆ , denoted as bladian superoperator, ~, is shown in Fig.1(c). The (`,`+1) (`,`+1) L Stot,Stot,z where Stot takes values 0, 1, and 2, zero eigenvalue is non-degenerate implying that the sin- | i(`,`+1) glet state is the unique steady state of the dynamics. and Stot,z takes all integer values in the range Stot Moreover, the distance of ρ (t) from the singlet state ob- Stot,z Stot. Additionally, we denote the AKLT− ground≤ s ≤ tained from the exact dynamics and measured via the state as ΨAKLT and the corresponding density matrix | i Frobenius or the trace norm (Eq. (18)) decays exponen- as ρAKLT = ΨAKLT ΨAKLT . tially in time with a rate given by ∆ δt which is in The AKLT| Hamiltoniani h with| periodic boundaries has excellent agreement to the numerically× obtained gap in a unique valence-bond ground state which is most easily the spectrum of the Lindbladian between the zero eigen- understood by noting that HAKLT can be expressed as a value and the rest of the eigenvalues. This shows that sum of local projectors the effective Lindblad equation is a valid description of N the dynamics. X (`,`+1) HAKLT = tot , (32) Note further that the form of the jump operators in PS =2 `=1 Eq. (29) leads the equation of the density matrix ele- ments in the basis spanned by the singlet and the triplets (`,`+1) where Stot=2 is a projector onto the five-dimensional to be of the form in Eq. (26). This explains the expo- Stot =P 2 subspace for the pair of spins-1 at sites ` and nential decay of the off-diagonal elements. Additionally, ` + 1. The form of HAKLT in Eq. (32) implies that the the master equation for the diagonal elements possesses ground state has the special property that each bond has only the loss term for the triplet states and only the gain zero weight in the Stot = 2 sector. Hence, it is natu- terms for the singlet, resulting the exponential decay and ral to imagine that a spin-1 chain can be steered to the growth of the former and latter respectively. AKLT ground state by locally steering each bond out of Before concluding the section, it is worth mentioning the Stot = 2 sector. Moreover, since the Stot = 2 sector that a protocol for steering a pair of spins-1/2 to the of each bond is finite-dimensional, such a steering pro- singlet using only a single detector qubit can also be en- tocol would satisfy our physically motivated constraints: visaged. Such a protocol involves steering the system out locality and only an extensive number of detectors. of any one of the triplet states onto the singlet state while In Sec.IVA we discuss the basic building block of the also acting on it with an intrinsic Hamiltonian which has steering protocol, namely steering a pair of spins-1 out matrix elements connecting the other two triplet states to of the Stot = 2 sector. In Sec.IVB we use this building the first one. Consequently, the Hamiltonian unitarily ro- block to steer a spin-1 chain to the AKLT state. tates the states within the triplet sector and weight from the triplet sector keeps leaking into the singlet, eventu- ally leading to the pure singlet state as the steady state. A. A pair of spins-1 For details, see AppendixC. Consider steering on the bond between a pair of spins- tot 1 labelled as ` and ` + 1. The S(`,`+1) = 2 subspace IV. SPIN-1 CHAIN is five-dimensional and hence the simplest protocol in- volves five detector qubits, one to steer out of each of tot Let us now discuss the steering of a quantum many- the S(`,`+1) = 2 states. Note that we only wish to steer tot body system, namely that of the spin-1 chain to the out of the S(`,`+1) = 2 subspace and there are no restric- AKLT ground state. The AKLT Hamiltonian is tot tions on what precise state(s) in the S(`,`+1) = 1 and Stot = 0 subspaces we steer onto. This offers a lot of N   (`,`+1) X 1 1 1 2 HAKLT = + Sˆ` Sˆ`+1 + (Sˆ` Sˆ`+1) , (30) freedom in choosing the steering protocol. This could be 3 2 · 6 · `=1 exploited to construct the simplest system-detector cou- pling Hamiltonian, or from a practical point of view, the where Sˆ` denotes spin-1 operators at site ` and the site one that it is easiest to implement. In the following, we ` = N + 1 is identified with ` = 1 to impose periodic choose a particular protocol which steers the state onto tot tot,z boundary conditions. the S(`,`+1) = 1 subspace keeping the sign of S(`,`+1) the It is useful to set up the notation for the rest of the same. section here. The projectors onto the three eigenstates At the start of each steering event, the detector ˆz ± qubits [41] are prepared in a polarised pure state of S` (with eigenvalues 1 and 0 respectively), P` and 0 ± P , are given by 5 z ` I2 + σ (`,`+1) (`,`+1) Y d ρ = i , (33) ± z 2 z 0 + − d P = [(S` ) S` ]/2; P` = I3 P P . (31) ⊗ 2 ` ± − ` − ` i=1 10 and the system-detector coupling Hamiltonian is of the where form

5 (`,`+1) X − (`,`+1) (`,`+1) Hs-d = J σ (`,`+1) ρd Ui + h.c., di ⊗ i=1 (34)

1 U (`,`+1) = ( 1, 1 2, 2 ) = [(S− S− )P +P + ], (35a) 1 | i h | (`,`+1) 2 ` − `+1 ` `+1 " − + ! + − ! # (`,`+1) 1 S` S`+1 + 0 S` S`+1 0 + U2 = ( 1, 1 2, 1 ) = I9 P P`+1 I9 P` P , (35b) | i h | (`,`+1) 2 − 2 ` − − 2 `+1

" − + 2 ! + − 2 ! (`,`+1) 1 (S` S`+1) + − (S` S`+1) − + U3 = ( 1, 0 2, 0 ) = I9 P P I9 P P | i h | (`,`+1) 2√3 − 4 ` `+1 − − 4 ` `+1 +(S+S− S−S+ )P 0P 0  , (35c) ` `+1 − ` `+1 ` `+1 " + − ! − + ! # (`,`+1) 1 S` S`+1 − 0 S` S`+1 0 − U4 = ( 1, 1 2, 1 ) = I9 P P`+1 I9 P` P , (35d) | − i h − | (`,`+1) 2 − 2 ` − − 2 `+1 1 U (`,`+1) = ( 1, 1 2, 2 ) = [(S+ S+ )P −P − ]. (35e) 5 | − i h − | (`,`+1) 2 ` − `+1 ` `+1

tot As in Sec.III, the jump operators for the corresponding support of ρs(t) on the S = 2 subspace decays expo- Lindblad equation in the time-continuum limit can be nentially, whereas it grows on the Stot = 1 and Stot = 0 identified straightforwardly as subspace before saturating to unity. An important point to notice is for a pair of spins-1, L(`,`+1) = JU (`,`+1)√δt. (36) i i the above dynamics does not have a unique steady state In the rest of this subsection, we only discuss the pair unlike the case exemplified in Sec.IIIC. This is due to of spins-1 and hence drop the labels ` and ` + 1. We also the fact that in the case of the spins-1, the steering is not refer to the density matrix of the two spins as ρs(t). The onto a particular pure state but rather onto a subspace dynamics of ρs(t) under the steering protocol with the of states, and the specific steady state of the dynamics system-detector coupling Hamiltonian in Eq. (34) start- depends on the initial condition of the system. The non- ing from a random mixed state is shown in Fig.2. Ex- uniqueness of the steady state for the pair of spins-1 is pressed in the Stot,Stot,z basis, the matrix elements of also manifested in the corresponding Lindbladian having tot| i ρs(t) in the S = 2 block and the off-diagonal blocks a degenerate zero-eigenvalue manifold. In the case of a connected to it decay exponentially in time. At long spin-1 chain, as we will discuss in the next subsection, enough times, ρs(t) is supported entirely on the subspace the uniqueness of the AKLT ground state leads to the spanned by the Stot = 1 and Stot = 0 states, and thus uniqueness of the steady state reached under dynamics is steered out of the Stot = 2 subspace. With regard obtained by extending the approach described above to to the corresponding Lindblad equation, note that the all the bonds of the chain. jump operators, Eq. (36), for the system-detector cou- pling Hamiltonian, Eq. (35), result in an equation of mo- tion of the form of Eq. (26) for the matrix elements of tot tot,z B. Steering to the AKLT state ρs(t) in the S ,S basis. Several features of the dynamics can| be inferredi from this observation. For any tot tot state α in the S = 1 or S = 0 sector, γνα = 0 Let us now discuss in detail the many-body case of a for ν =| αi which renders the off-diagonal matrix elements chain of spins-1. The protocol for steering the system within6 the Stot = 1 and Stot = 0 subspace invariant in to the AKLT ground state is a scaled-up version of the time (see also Fig.2(a1)-(a4)). This also implies that protocol described in the previous subsection where each the master equations for the diagonal elements within bond is steered out of the Stot = 2 subspace and as such this subspace have no loss terms and only gain terms, there are 5N detector qubits at play, N being the size contrary to those in the Stot = 2 subspace, which only of the spin-1 chain. Formally, the initial state of the have loss terms. One can then immediately infer that the detectors at the start of each steering event is expressed 11

− Jt = 0.1 Jt = 1.0 Jt = 10.0 Jt = 100.0 is of the form σd 1, s1 2, s2 + h.c.. Hence a state (a1) (a2) (a3) (a4) 1 ⊗ | i h | of the form ρd ρAKLT is annihilated by the Hamilto- Stot = 2 + ⊗ nian as σ ρd = 0 = 1, s1 2, s2 ρAKLT, implying that d | i h | ρAKLT is indeed a steady state of the dynamics. How- Stot = 1 ever, compared to the ‘ideal’ steering protocol described Stot = 0 0 in Sec.III, a number of difficulties arise for many-body = 2 = 1 = 0 = 2 = 1 = 0 = 2 = 1 = 0 = 2 = 1 = 0 systems. The difficulties arise because the Hilbert space tot tot tot tot tot tot tot tot tot tot tot tot S S S S S S S S S S S S for a many-body system is exponentially large in system i (b) ,z 2, 2 2, 0 2, 2 1, 0 0, 0 size. The protocol of Sec.III uses a separate term in H tot | i | i | − i | i | i s-d 2, 1 2, 1 1, 1 1, 1 ,S 0.4 | i | − i | i | − i to steer to the target state from each other state. Such an tot S

| approach is not practical for a many-body system, where s ρ |

,z one requires a limit of at most an extensive number of 0.2 tot terms in Hs-d. In the following we discuss these issues, ,S

tot and present numerical results for the AKLT chain which S

h 0.0 1 0 1 2 show successful steering in this many-body setting. 10− 10 10 10 Jt In the derivation of the strong inequality for steering in the presence of multiple detectors (see Sec.IIC and FIG. 2. Steering of a pair of spins-1 out of the Stot = 2 (m) (n)† AppendixB), we had required that Us Us = 0 if subspace using the system-detector coupling Hamiltonian of (m) (n) Us and Us have overlapping spatial supports. For the Eq. (34). (a) The density matrix of the system, ρs in the |Stot,Stot,zi basis as a colourmap (for the absolute values of system-detector couplings with Eq. (35) for the AKLT (`,`+1) (r,r+1)† the matrix elements) at different times t: it shows the decay- chain, it can be shown that Ui Uj = 0 if tot tot 6 ing and growing support on the S = 2 and the S = 0 and ` r = 1. In other words, the assumption made in the 1 subspaces respectively. The blocks demarcated by the blue, | − | tot derivation is not satisfied by the system-detector cou- red, and gray dashed lines respectively denote the S = 2, 1, plings on adjacent bonds but it is satisfied by all the and 0 sectors. (b) The evolution of the diagonal elements of tot other pairs of couplings. The non-commutativity of the ρs(t) shows the steering out of the S = 2 subspace, all the five diagonal elements of which decay to zero. For the plots, steering operators on adjacent bonds may also lead to J = 1 and δt = 0.1. non-monotonic time-dependence for EAKLT(t). Steering a bond, say between sites ` and ` + 1, affects the state of three bonds: the one that is steered and the two adjacent as to it. While the energy of the bond being steered neces- h (`,`+1) i sarily goes down as steering decreases Tr Stot=2 ρs(t) , z P N 5 2 + σ (`,`+1) h 0 0 i Y Y I d (` ,` +1) i the same need to be the case for Tr tot ρs(t) with ρd = , (37) S =2 ⊗ 2 0 P `=1 i=1 ` = ` 1. For this reason, the total energy of the state need not± decrease. and the system-detector coupling Hamiltonian is of the The difficulty associated with the non-commutativity form of the steering operators can be partially removed by con- sidering a protocol in which each steering step is divided N 5   into multiple steps, in one of which only the system- X X − (`,`+1) Hs-d = J σ (`,`+1) ρd Ui + h.c., (38) detector couplings on a particular bond acts. In this di ⊗ `=1 i=1 case, at any time only one bond is steered. For the op- (`,`+1) (`,`+1)† erators acting on the same bond, Ui Uj = 0 (`,`+1) tot where Ui is the same as in Eq. (35). as the different S = 2 states on each bond are or- Let us first argue that the protocol described by thogonal to each other. Hence, the inequality is satisfied Eqs. (37) and (38) does posses the AKLT ground state in each of the steps. The evolution of the diagonal el- at least as one of its steady states. Note that each term ement of the system’s density matrix corresponding to in the system-detector coupling Hamiltonian in Eq. (38) Ψs = ΨAKLT is then described by | ⊕ i | i

5 X (`,`+1) (`,`+1)† 2 Ψs ρs(t + (` + 1)δt) Ψs = Ψs ρs(t + `δt) Ψs + Ψs U ρs(t + `δt)U Ψs sin δt, (39) h ⊕ | | ⊕ i h ⊕ | | ⊕ i h ⊕ | i i | ⊕ i i=1 | {z } Q`

where the equation corresponds to the specific case of steering the bond between sites ` and ` + 1. The sum 12

Jt = 0.1 Jt = 1.0 Jt = 10.0 Jt = 100.0 on the right-hand side of Eq. (39) is over all the detector 1 qubits acting only on the bond between sites ` and ` + 1. (a1) (a2) (a3) (a4) Stot = 2 Since U (`,`+1)U (`,`+1)† = 0, Eq. (39) follows from the i j 6 derivation of Eq. (B5) (AppendixB). Note that Q` 0 tot ≥ S = 1 for all ` as for Q in Eq. (16). In the time-continuum limit tot S = 0 0 δt 0, the set of equations in (39) for all ` give → = 2 = 1 = 0 = 2 = 1 = 0 = 2 = 1 = 0 = 2 = 1 = 0 tot tot tot tot tot tot tot tot tot tot tot tot

2 S S S S S S S S S S S S

Ψs⊕ ρs(t + Nδt) Ψs⊕ = Ψs⊕ ρs(t) Ψs⊕ + (δt) Q, i ,z h | | i h | | i 2, 2 2, 1 1, 0 i 0 D1 (40) tot 10 | i | − i | i D P ,S 2, 1 2, 2 1, 1 F with Q = Q`. 0.4 AKLT ` tot | i | − i | − i ∆t Ψ

| 1 e S 2, 0 1, 1 0, 0 −

| 10 Since Q 0, Eq. (40) ensures that the state is never | i | i | i − ∼ (2) s

≥ ρ steered further away from the target AKLT state if Q is | 2 ,z 0.2 10− employed as a measure of distance from the target state. tot ,S 3 tot At the same time Q = 0 does not necessarily ensure (b) distance from 10− (c) S 0.0 h 1 0 1 2 that the system is in the target state. As an illustra- 10− 10 10 10 0 200 tion, consider an initial pure state that has excitations Jt Jt on multiple bonds. A single steering step of the type we have described removes one excitation without generat- FIG. 3. Steering of a spin-1 chain to the AKLT ground ing overlap with the target state, and so Q = 0 for this state. (a) The evolution of the reduced density matrix of (2) step. So while the state changes, its overlap with the two adjacent spins, ρs , shown as a colourmap (for the abso- lute values of the matrix elements) in the |Stot,Stot,zi basis. AKLT state does not. tot In fact, for our steering protocol it can be shown ex- The support on S = 2 sector decays to zero and so do the off-diagonal elements in this basis. (b) The evolution of the plicitly that there can arise ρs(t) for which the change (2) diagonal elements of ρs . The horizontal dashed lines denote in EAKLT(t) (defined in Eq. (19)) is positive. Using the values 1/3 and 2/9, obtained from ρ(2) . (c) The dis- EAKLT(t) as a measure of distance one would then con- AKLT clude that the state is steered further away from the tance between ρs(t) and ρAKLT as measured via the matrix norms D and D decays exponentially to zero with time. AKLT ground state. To show this we use the fact that F 1 The dashed line represents an exponential decay with a rate steering a bond affects only three bonds or equivalently ∆ obtained from the spectral gap of the corresponding Lind- the four sites involving them. Since HAKLT is a sum of bladian. For the plots, N = 5, J = 1, and δt = 0.1. projectors for each bond, the change in energy can be expressed to leading order in δt as (see AppendixD for (2) details) spins-1 which we denote as ρs , and (ii) the distance of h i ρs(t) from ρAKLT as a function of time, measured via D1 2 ˆ ∆EAKLT(t) = (δt) Tr `ρs(t) , (41) and DF . − E (2) The results are shown in Fig.3. The evolution of ρs (t) tot where ˆ` is a four-site operator involving sites ` shows that each bond is indeed steered out of the S = 2 E − 1, , ` + 2. Since the spectrum of the operator ˆ` subspace. However, a drastic difference between the dy- ··· E (2) includes eigenvalues of both signs, monotonic decay of namics of ρs and that of a single pair of spins-1 is that EAKLT(t) is not guaranteed. On the other hand, we find the former has a unique steady state which is diagonal that for a choice of ρs(t) which leads to an initial growth tot tot,z (2) in the S ,S basis with 1, s1 ρs (t ) 1, s1 = of ˆ , the distance from the target measured via D (t) | (2)i h | → ∞ | i ` F 2/9 and 0, 0 ρ (t ) 0, 0 = 1/3. This highlights or ED (t) nevertheless decays monotonically. A detailed s 1 the collectiveh | effect→ of the∞ | many-bodyi yet local steering study of the monotonicity of steering, or lack thereof, protocol, as each of the system-detector couplings are depending on the distance measure, is presented in Ap- such that they keep the off-diagonal elements within the pendixD. The upshot of this study is that one needs to Stot = 1 and 0 subspace invariant. Moreover, note from show that the target state is the unique steady state of the terms in Eq. (35) that they do not change the di- the dynamics. As any state of the spin-1 chain which agonal element corresponding Stot = 0. On the other is not the AKLT state will have some overlap on the (2) Stot = 2 sector on one or more bond, the form of the hand, in the many-body case, the matrix element of ρs operators in Eq. (35) ensures that such a state is not a on a particular bond goes to its steady state value of 1/3 stationary state. purely due to the interplay of the steering on the bond To demonstrate that the AKLT state is indeed the with that on the other bonds, further highlighting the collective effects of the many-body steering protocol. unique state and also to show the validity of results away (2) from the time-continuum limit, we simulate numerically The fact that ρs approaches a diagonal form in the tot tot,z the dynamics of a spin-1 chain starting from a random S ,S basis with the diagonal elements taking the | i (2) mixed state subjected to the steering protocol with all the above mentioned specific values is crucial because ρAKLT bonds steered simultaneously. We calculate two quanti- has exactly this feature. For the reduced density ma- ties: (i) the reduced density matrix of a pair of adjacent trix of a subsystem of the spin-1 chain comprising the 13

(a) 0 (b) ED of L it most convenient to answer these questions in terms of approach to SS ) 0.4 the Lindbladian corresponding to the steering protocol. i| 2 − Following the discussion in Sec.IIIB, the jump operators,

AKLT 0.3

Ψ of which there 5 N, can be immediately identified as | )

t 4 N = 5 (`,`+1) (`,`×+1) ( − ∆ √ ψ N = 6 0.2 Li = JUi δt. For convenience, we will set

− |h 6 N = 7 J√δt = 1 in the following. As discussed in Sec.IIIB, − N = 8 0.1 the uniqueness of the steady state manifests~ itself in the log(1 8 N = 9 non-degeneracy of the zero eigenvalue of ~. The rate of − N = 10 0.0 the approach to the steady state is given byL the gap ∆ of 0 10 20 0.0 0.1 0.2

Jt the rest of the spectrum from the zero eigenvalue. Hence, 1/N ~ it is instructive to study the scaling of ∆ with N. By exactly diagonalising ~ for finite sizes we find that FIG. 4. Scaling of steering rate with system size for L

the spin-1 chain. (a) The exponential approach of the spin- the zero energy eigenvalue is indeed non-degenerate and 1 chain to the AKLT state as obtained from the wave-function the unique steady state corresponding~ to it is identical to 2N Monte Carlo method for solving the Lindblad dynamics for ρAKLT. The dimension of ~ is 3 , which restricts calcu- different system sizes N. As the method involves evolving lations to N 8. In orderL to extract ∆ for slightly larger pure states, we measure the distance from the AKLT state sizes, N 10,≤ we solve for the dynamics of the Lindblad via log(1 − hψ(t)|ΨAKLTi) where the bar denotes the average equation≤ using the wave-function Monte Carlo method over the wave-function trajectories. The black dashed lines which involves evolving pure states with stochastic quan- are fits to the data used to extract ∆. (b) The rate of the ex- tum jumps and then averaging over the so-obtained tra- ponential approach to the steady state (SS) obtained from (a) or equivalently the gap in the spectrum of the Lindbladian ob- jectories [43, 44]. With this method, we find again that tained from exact diagonalisation of L shown as a function of the spin-1 chain approaches the AKLT ground state ex- 1/N. The intercept indicates a finite ∆ in the thermodynamic ponentially fast in time and hence the rate ∆ can be

limit. The data for panel (a) was averaged over 20000 trajec- extracted, see Fig.4(a). The ∆ obtained from the dy- tories and the errorbars were obtained via standard bootstrap namics for smaller system sizes agrees~ very well with the with 500 resamplings. Errors in the extrapolated value of ∆ one obtained from diagonalising ~, see Fig.4(b). The introduced via the errors in ∆ for different N as well as the two methods together allow us toL study the scaling of fitting procedure are also shown but the error bars are smaller ∆ with N and from the scaling shown in Fig.4(b), we than the point size. conclude

∆(N) = ∆∞ + c/N . (42) contiguous sites `1 through `n, it can be easily argued ~ that its eigenvectors corresponding to non-zero eigenval- Hence the gap in the the spectrum of ~ or equivalently ues are the four degenerate AKLT ground states of the the time-scale required to steer the spin-1L chain to the spin-1 chain of length n with open boundary conditions AKLT state stays finite in the thermodynamic limit. P`n (`,`+1) described the Hamiltonian Hsubsys = tot . At Combining all the results presented in this section, we `=`1 PS =2 P`n 2 conclude that the map of Eq. (2), with the detectors’ the same time, Hsubsys commutes with ( Sˆ`) and `=`1 initial states and the system-detector coupling described P`n Sˆz. These eigenstates are simultaneous eigen- `=`1 ` by Eqs. (37) and (38) respectively, constitutes a protocol P`n Sˆ 2 P`n ˆz states of Hsubsys,( `=`1 `) , and `=`1 S` . For n = 2, for steering a many-body system to a strongly correlated (2) tot tot,z it directly implies that ρs is diagonal in the S ,S state, in this case a spin-1 chain to the AKLT ground basis. Moreover, it has been shown that the four| non-zeroi state. The protocol satisfies the constraints of having an eigenvalues are (1+3( 3)−n)/4 and (1 ( 3)−n)/4 with extensive number of detector qubits coupled only locally a threefold degeneracy− [42] which for n−=− 2 are 1/3 and to the system and of having a steering time-scale that 2/9 respectively. Since every bond of the spin-1 chain is does not diverge in the thermodynamic limit. steered to a state that corresponds to the AKLT ground state of the entire chain, it is natural that the many-body state ρs(t) is also steered to ρAKLT. This confirmed by V. DISCUSSION the exponential decay of the distances, DF and D1 be- tween ρs(t) and ρAKLT with time as shown in Fig.3(c). In this work we have developed a general protocol that So far we have argued that ρAKLT is a steady state of uses entanglement to induce steering of a quantum sys- the steering protocol and have shown numerically that tem to a target state. We have shown how our protocol an arbitrary initial state of a finite spin-1 chain evolved can be applied both to steer systems with a few degrees with the steering protocol approaches the AKLT ground of freedom and to prepare strongly correlated states in state exponentially fast in time. The natural question to many-body systems. The form of steering that we study ask then is what happens to the steering protocol in the here is rooted in but distinct from steering of the type dis- N limit, with regard to both the uniqueness of the cussed originally in the context of the EPR paradox [1,2]. steady→ ∞ state and the time-scale for steering to it. We find Specifically, our protocol shares with earlier discussions 14 of quantum steering the fact that it exploits the entan- not being in the third level. This can be represented us- glement between the system of interest and another sys- ing modified Lindblad dynamics leading to non-hermitian tem. Crucially, however, it differs in offering a general Hamiltonian dynamics. Alternatively, the dynamics can framework to steer a quantum system into a specific and be described in terms of a set of null weak measurements pre-designated state. with post-selected readout sequence [51, 52]. Either pic- ture differs substantially from our protocol. However, ad- justing these protocols to our paradigm of steering should A. Experimental realisations be rather straightforward. Generalizations of these pro- tocols could work for a detector coupled to two or more qubits [53]. Experimental implementations of the ideas presented in this work require realisations of qubits which satisfy The second possible realization of superconducting several conditions. (i) The relaxation time of the qubit qubits involves the use of state-of-the-art fluxonium due to undesired coupling to external degrees of freedom qubits. These offer coherence times comparable to more should be long. (ii) Likewise, the coherence times should common qubits, with increased anharmonic- be long. (iii) It should be possible to design system- ity, reduced cross-talk, and design flexibility. Such de- detector couplings as required for versatile steering to vices could be implemented to construct both our many- arbitrary target states. (iv) The system should be scal- body system as well as the detector qubits [54–57]. Con- able to multiple qubits and should allow for multi-qubit cretely, two fluxonium qubits could be coupled with vari- interactions (for example a three qubit-interaction term able strengths. One fluxonium plays the role of the qubit in the system-detector Hamiltonian involving one detec- to be steered, while the other fluxonium, inductively cou- tor qubit and two system qubits). pled to a readout resonator [56], plays the role of detector There are several promising candidate platforms for on which dispersive projective readout can be performed realising our protocol, which include superconducting with high fidelity. The coupling between the qubits can qubits, cold ions, nitrogen-vacancy (NV) centres, macro- be implemented using shared capacitors or shared induc- scopic mechanical objects with entangled quantum de- tors. With the recent availability of compact inductors grees of freedom, and systems. These with low nonlinearity, using disordered superconductors are platforms that have all been tested, or will be tested, such as granular aluminum [58] or TiN [59], inductive as far as the steering of a single qubit is concerned. The coupling appears more appealing, as it minimizes cross- extent to which more complex systems, comprising mul- talk between the qubit to be steered and the readout tiple quantum degrees-of-freedom, may be steered relies resonator. on the extent to which conditions listed earlier can be sat- Cold ions: Another prospective platform for imple- isfied. In the following we briefly discuss some of these menting measurement-induced steering is based on cold platforms. ions trapped by magnetic field and lasers. Qubits are Superconducting qubits: The most prospective experi- implemented in electronic states of each ion and the ions mental avenue for our proposal is based on superconduct- in a common trap can communicate with each other ing qubits. There are two realizations of qubits which ap- via intrinsic long-range interactions and external, laser- pear to be the leading candidates for constructing steer- induced couplings. Below, we will describe an example + ing platforms. The first are . In previous ex- of such a platform consisting of Sr ions which are es- perimental work, weak measurements were constructed sentially hydrogen-like atoms with one outer electron; the via entangling operations between two superconducting lowest two levels represent the qubit. Coupling to the ion qubits, and subsequent projective measurement on the with light can effect arbitrary unitary rotations. Qubit detector qubit [45] which is exactly the measurement operations are well-developed in this area, but steering structure we propose in this work. Another approach, im- of a single qubit is only underway, while a challenge for plemented primarily for the steering of a single qubit, re- applications to steering is to develop multiple-qubit op- lies on dispersive interaction between the qubit and a cou- erations. pling to a waveguide cavity. Existing implementations of Generalizing from single-qubit to two-qubit measure- weak measurement back-action on such a systems [46–49] ments is highly challenging. The more straightforward have three major differences from our steering protocol. protocol would involve shining light of a wavelength First, the sequence of weak measurements in the experi- larger than the distance between ions, so that each ion ments amounts to a strong measurement when integrated can absorb and emit a photon, ensuring coherence be- over time (i.e., when considering a sequence of many weak tween the two ions [60–63]. The detection of an emitted measurements), which projects the initial state onto a photon in a given direction determines the backaction. final state with probabilistic outcomes. Second, the ex- To achieve versatility of the system-detector Hamilto- periments make use of post-selected readouts. Third, the nian, one needs to go one step further and designate one specifics of the experiment rely on knowledge of the sys- ion as a detector that measures two other ions (one on tems initial state. A further experiment [32, 50] focusses the left and the other on the right of it). Every two neigh- on the study of a three level system, where the evolution bouring ions can be made to interact through terms of the of a two-level subsystem is conditioned on the system system-detector Hamiltonian that flip the pair of qubits. 15

If it is possible to couple and decouple on-demand the de- that other types of manipulations are, in principle, also tector and the left spin, and separately the detector and possible. Going to the arena of many-body physics, the right spin, then performing unitary transformations one needs to engineer scalable multi-qubit states by our on the right spin-detector and/or the left spin-detector, steering protocols. This is feasible with platforms based a broad class of effective 3-qubit Hamiltonians can be on superconducting qubits and NV centres, where flex- achieved. ible (and, in principle, arbitrarily complex) few-qubit- Nitrogen-vacancy (NV) centres: NV centres in dia- detector couplings can be engineered, as demonstrated monds are point defects which realise electronic spins lo- for two-qubit realizations. Implementation of many-body calised at atomic scales. In these setups, the detector steering on these platforms requires experimental efforts could be the NV centre and the qubits could be 13C nuclei towards long coherence times, circuit connectivity, and spins. The interaction between them is of a dipole-dipole quality control of system-detector coupling and detec- type, which may be controlled with high precision [64]. tor re-initaition. Importantly, even when some specific Direct optical measurement can only be preformed on the system-detector couplings are still hard to design for NV center. The control is made with microwaves [65]. technological reasons, the shake-and-steer protocol may Another possibility is to use NV centres to control exter- resolve this issue. We expect our work to stimulate ex- nal spins [66]. In principle, one could aim at a multi-spin periments based on existing and future platforms in the measurement or control. The level structure for such direction of steering and manipulating many-body states. systems has been analysed extensively [67]. Various as- pects of the NV readout and polarization mechanism [68] as well as control at weak coupling have been studied B. Connection to previous work (cf. Refs. [69] and [70] for theory and experiment respec- tively). We note that the control can be much improved The use of external degrees of freedom as a designed at low temperatures [71]. environment for manipulating and controlling quantum Macroscopic mechanical objects: Measuring and con- states has been discussed previously in other settings [23, trolling macroscopic mechanical objects that exhibit 29, 30, 79]. Treatment of the environment as Markovian quantum coherence is a challenge realised in recent or the measurements as non-selective leads to a Lind- years [72]. The next step, accomplished very re- blad equation, similar to one derived in the present work. cently [73], is to scale this up to two macroscopic me- Much of the earlier focus has been either on the formal chanical degrees of freedom. Unfortunately, achieving a aspects of such non-unitary dynamics, or on applications high level of control and measurement may expose the involving few-body quantum systems. Our work consti- system to undesired interactions with its environment, a tutes a generalisation of such quantum control and steer- problem that becomes more pronounced at the macro- ing to macroscopic many-body systems. scopic scale. Using a superconducting electromechani- In the context of open quantum systems, the cre- cal circuit and a pulsed microwave protocol, a system of ation of non-trivially correlated or topological many- two mechanical drumheads has been steered towards an body states has been proposed using so called drive- entangled state, where entanglement of continuous de- and-dissipation schemes [17–19, 22, 24–26, 80]. In such grees of freedom satisfies the criteria formulated [74, 75]. schemes the system is driven or excited using a time- The strength of the measurement in these systems can dependent Hamiltonian while a dissipative channel, of- be tuned from strong to weak, and coherence may be ten a Markovian environment, working simultaneously maintained over long time-scales. However, given the relaxes the system. This interplay of drive and dissipa- immense experimental difficulties here, the issues of ver- tion is then used to engineer non-trivial states. From a satility of system-detector Hamiltonian, and the question theoretical point of view, the Markovian nature of the of scalability to systems consisting of multiple degrees of environment naturally lends itself to a treatment via the freedom are all likely to remain a challenge in the near Lindblad equation, hence yielding a formal connection to future. our measurement protocol in the time-continuum limit Finally, one could adopt these ideas to the field of alluded to previously. In fact, the jump operators postu- quantum optics [76]. Photon polarization has been used lated in Ref. [18] in the context of the AKLT chain also as a qubit, for example, in the process of storing and bear formal similarities with those derived in our case then retrieving quantum information in cold atom plat- from the measurement protocol. forms [77]. More directly related to our ideas are two Further related work is described in Ref. [20], where polarization-based qubits (one serving as system and the auxiliary qubits are exploited to drive a Rydberg-atom other as detector) allowing for back-action control [78]. realisation of Kitaev’s toric code to its ground state. Ex- In summary, experimental implementations of our cited states of the toric code contain spatially localised ideas should be assessed according to the complexity of excitations on plaquettes or vertices of the system, which the required platform. As far as a single-qubit steer- can annihilate in pairs. The protocol of Ref. [20] takes ing is concerned, reliable implementations are achiev- advantage of this special feature of the toric code. To able with all platforms. Entangling two qubits has al- cool this system to its ground state, an implementation ready been done with some platforms, with the promise of Lindblad jump operators is proposed that generates 16 stochastic motion of excitations, so that pairs meet and Another direction for taking non-Markovian effects into annihilate. An advantage of our steering protocol for the account is via the quantum collision models [81]. We re- AKLT chain is that we directly steer each local degree serve the generalisation of our protocol to include such of freedom of the system (a bond between two sites in effects for future work. the AKLT chain) towards the ground state. Moreover, There are many possible motivations for such general- our general protocol is independent of special features isations. One would be to optimise the time-scale over of the steered system. The local unit which is steered which the system reaches a pre-defined vicinity of the could have a large excited state manifold; steering would target state. Another would be to maximise the purity still be possible using the shake-and-steer protocol (see of the quantum state of the system throughout the pro- AppendixC) with a finite number of locally coupled de- tocol. This would become particularly important if one tectors. extends the objective of the steering protocol from prepa- In contrast to the drive-and-dissipation protocols, our ration of a target state to manipulation of a state using measurement-based protocol does not require a system a sequence of weak measurements [82]. Hamiltonian to induce dynamics in the system. More- Another set of open questions concerns the robustness over, unlike an environment which is to a large extent un- of the protocol to deformations of the initial states of controlled, the detectors that play the role of the dissipa- the detector qubits or of the system-detector couplings. tive channel are well-controlled simple quantum systems These questions call for a detailed and comprehensive that can be tuned in time. As on-demand jump opera- study of the consequences of errors in steering protocols tors, they represent a particularly a useful form of quan- and possible ways of correcting them, which we leave for tum manipulation and control. Moreover, as outlined a future work. Specifically, one could ask under these above, using the measurement outcomes for active de- conditions whether the system has a steady state, and if cision making, or having a measurement protocol which so how its distance from the target state scales with the varies with time and has a memory kernel, takes us to strength and probability of errors in the steering protocol. realms that seemingly cannot be addressed by drive-and- Furthermore, one might look for specific error correction dissipation techniques. protocols akin to a stabiliser formalism [83], but within the setting of a steering protocol. It is reasonable to anticipate that a finite gap in the C. Future directions spectrum of the Linbladian in the thermodynamic limit (as we find in our study of the AKLT state) will en- There is a wide range of possible future directions and sure that the steady state is robust to small random er- open questions arising from the work presented here, and rors in the steering protocol that are local in space and we close by outlining some of these. While the protocol time. Robustness of the protocol to time-independent we have described is arguably the simplest version, it deformations is a separate issue. A specific instance con- is also quite robust in the sense that the probability of cerns the interplay of a system Hamiltonian (which we reaching the target state is unity irrespective of the ini- have so far omitted) with the system-detector coupling tial state of the system. One can also envisage a wide Hamiltonian. This would amount to having a non-zero variety of generalisations, in which instead of simply de- V in Eq. (23). It is clear that if the Lindbladian in the coupling the detector qubits from the system of interest absence of the system Hamiltonian shares an invariant at the end of each steering step, a projective measure- subspace with the system Hamiltonian, the steering pro- ment is made of the final detector states. In this broader tocol is robust to the latter. However, if that is not the context, the protocol we have described corresponds to case, the apparent tension between the system Hamil- the special case of blind measurements in which an un- tonian and the system-detector coupling Hamiltonians biased average is made over all such measurement out- could lead to a new steady state which deviates continu- comes. A natural direction is to ask whether making use ously with the strength of the system Hamiltonian from of the measurement outcomes can allow optimisation of the original dark state. Alternatively, in the many-body the protocol. More precisely, the initial state of the detec- case there could be a dynamical phase transition in the tor, the system-detector coupling, and the duration of the nature of the steady state. coupling at each step can be made to depend on the mea- In fact, measurements on an otherwise random uni- surement outcome of the previous step or on an extended tary circuit have been shown to induce an entanglement history of the measurement outcomes. Active error cor- phase transition in the steady state as a function of the rection could be viewed as an example of a broader class density or strengths of measurements [84–93]. Most of of protocols involving such decision making. Generalised these works use local projective measurements or generic protocols of this type are manifestly not Markovian and positive operator-valued measurements which manifestly hence are not describable in terms of a Lindblad equation decrease the entanglement. However, our work shows with time-independent jump operators. In the context of that measurement protocols can be constructive in the quantum control [79], there have been proposals to utilise sense that they can be designed so that steady state is a the measurement outcomes as a feedback to the dynamics strongly correlated state. Thus a natural question arises using the so-called quantum filtering equations [29, 31]. as to the extent to which measurement protocols and 17

Grant VP1-2015-005, and from the Deutsche Forschungs- 0.8 gemeinschaft (DFG) Projektnummer 277101999, TRR 183 (project C01), and Grants No. EG 96/13-1 and GO 0.6 1405/6-1. 0.4

0.2 s ρs s Appendix A: Steering a single spin-1/2 to an h↑ | | ↑ i s ρs s arbitrary state 0.0 h↑ | | ↓ i 100 101 Jt In the toy example of a single spin-1/2 presented in Sec.IID, the target state of the system qubit was the same as the initial state of the detector. In this appendix, FIG. 5. Steering of a single spin-1/2 using the system detec- we demonstrate that this is not a requirement, and show tor coupling in Eq. (A3) with θ = π/4. The different lines show the different matrix elements of the system spin’s den- that our protocol can be used to steer the system qubit to sity matrix. The horizontal dashed lines show the values that an arbitrary target state. Let us suppose that we always prepare the detector spin in its d state, but we want the matrix elements should take for ρs = |Ψs⊕ i hΨs⊕ |. to steer the system spin to a pure|↑ statei of the form

Ψs = cos(θ/2) s + sin(θ/2) s . (A1) their interplay with system Hamiltonians can be classi- | ⊕ i |↑ i |↓ i fied in terms of entanglement properties of the resulting This is easily achieved using a system detector coupling steady states. Hamiltonian of the form An intriguing future direction would be to study mea- − y surement protocols for which the emergent Lindbladian Hs-d = Jσ Ψs⊕ Ψs⊕ ( iσs ) + h.c., (A2) d ⊗ | i h | − has a dark space spanned by several dark states. An ex- y where iσ Ψs is the state orthogonal to Ψs and out ample is provided by the AKLT state in an open chain, − s | ⊕ i | ⊕ i for which this space is four dimensional. In such a sce- of which we want to steer. In terms of the spin operators, nario one can envision a closed adiabatic trajectory (in the Hamiltonian can be written as the protocol parameter space) that could be used to in- J z x y − duce a unitary transformation in the dark space. This Hs-d = (sin θ σ cos θ σ iσ ) σ + h.c. (A3) 2 s − s − s ⊗ d could be harnessed to induce adiabatic rotations in a de- generate many-body space, and may even give rise to a The resulting dynamics of ρs under the protocol with the many-body non-Abelian geometric phase [94, 95]. Simi- system-detector coupling Hamiltonian as above is shown larly, the presence of multiple stationary states of Lind- in Fig.5 for θ = π/4. The different lines correspond bladians has been used theoretically to realise dissipative to the matrix elements of the time-evolving density ma- 2 time-crystals [96–99], and these could in principle also be trix of the spin with ρs(t ) = cos (θ/2) and h↑ | → ∞ | ↑i generated via a measurement protocol. ρs(t ) = cos(θ/2) sin(θ/2) at long times show- h↑ | → ∞ | ↓i ing that in the steady state ρs Ψs Ψs . → | ⊕ i h ⊕ |

ACKNOWLEDGMENTS Appendix B: Derivation of steering inequality with multiple detector qubits E. Karimi, K. Murch, R. Ozeri, I. Pop, A. Retzker, and K. Snizhko have made insightful comments on ex- In this appendix, we present some detail of the deriva- perimental realisations. We have benefited from use- tion of the steering inequality in the presence of multiple ful discussions with Y. Herasymenko, P. Kumar, and K. detectors. Partitioning the composite Hilbert space of Snizhko. We acknowledge financial support from EPSRC the system and the detectors into ρd s and ( d) s Grant Nos. EP/N01930X/1 and EP/S020527/1, from as in Sec.IIIA, the system-detector⊗ coupling H Hamilto-D ⊗ H the Israel Science Foundation, from the Leverhulme Trust nian of Eq. (7) can be expressed as 18

(1) (n) Φd O Φd O Φd | i · · · d | i · · · d | i · · ·  † †  Φd 0 0 U1 Un .| i  ··· ···  .   . 0 0 0   . ·········. . . .  O(1) Φ  U ......  d d  1  Hs-d = . | i  ...... , (B1) .  ......  .  . . .  (n)  . . . . .  O Φ  U ......  d d  n  . | i ...... 0 ...... 0

(n) where the labels Od Φd on the rows and columns of the matrix denote the positions of non-zero entries, Un and † { | i} th Un. The entry labelled by n corresponds to the coupling of the system with the n detector. With this notation, the operator W = exp ( iHs-dδt) takes the form −  Xk −iδt Xk[ U † U † ] 2k+1 ··· 1 ··· n ···   0   ∞   ( iδt)2k  U1   X   .   W = −  −iδt k  .  k  , (B2) (2k)!  Y  .  Y  k=0  2k+1     Un    .   . where  .  .   U1  X †  .  † † X = U Un; Y =  .  [ U U ]. (B3) n  .  ⊗ ··· 1 ··· n ··· i   Un  .  .

Using the form of the operator W from Eq. (B1) in the map described by Eqs. (1) and (2), we obtain the time-evolved density matrix of the system as

∞ 2k 2l ∞ 2k+1 2l+1 X ( iδt) (iδt) k l X ( iδt) (iδt) X k l † ρs(t + δt) = − X ρs(t)X + − UnX ρs(t)X U . (B4) (2k)!(2l)! (2k + 1)!(2l + 1)! n k,l=0 k,l=0 n

The quantity of interest is Ψs⊕ ρs(t) Ψs⊕ , which can be obtained from Eq. (B4) under the condition that either (m) (n)† h (m|) | (n)i† Us Us = 0 for n = m or Us and Us commute, in the form 6 X (n) (n)† 2 Ψs⊕ ρs(t + δt) Ψs⊕ = Ψs⊕ ρs(t) Ψs⊕ + Ψs⊕ Us ρs(t)Us Ψs⊕ sin (δt), (B5) h | | i h | | i n h | | i

which is the same as equation as Eq. (16) with Q given one system-detector coupling. An additional Hamilto- by Eq. (17). This concludes our derivation of the steering nian acting on the system unitarily rotates the state of inequality for the case of multiple detectors. the system within the orthogonal subspace. Thus the weight from the orthogonal subspace keeps leaking into the target state via one channel corresponding to the Appendix C: Shake-and-steer protocol system-detector coupling, while the weight from the rest of the orthogonal subspace is shaken into the dissipative In all the previously discussed cases, the number of channel via the system Hamiltonian. Let us illustrate the system-detector couplings used locally is equal to the di- protocol using the steering of two spins-1/2 to the singlet mension of the subspace orthogonal to the target state. state, similar to the one in Sec.IIIC. The shake-and-steer protocol permits steering with just Out of the three system-detector couplings used there, 19

Appendix D: Monotonicity of steering to the AKLT 1.0 (a) (b) state 0.8 2 S ρ S h 0| s| 0i 0.6 T ρ T ] The numerical results in Sec.IVB show that the AKLT

s v h −| | −i 0

T ρ T Im[ ground state is the unique steady state of our measure- 0.4 h 0| s| 0i T ρ T ment dynamics and steering to it is guaranteed via our h +| s| +i 0.2 2 − protocol. However, the approach to the steady state is 0.0 not necessarily monotonic and can depend on the dis- 100 101 102 103 0.3 0.2 0.1 0.0 tance measure used. In this section, we show explicitly Jt − −Re[v] − that such a situation arises using EAKLT [Eq. (19)] as the distance measure. We then compare the results with those of other distance measures. FIG. 6. Shake-and-steer protocol for steering a pair P (`,`+1) Since HAKLT = tot is a sum of projectors for of spins-1/2 to the singlet state. (a) The dynamics of ` PS =2 the diagonal elements of the density matrix under the pro- each bond, we would like to understand the change in (`0,`0+1) tocol described in Eq. (C3) with θ1 = θ2 = J(1, 1, 1). (b) Tr[ Stot=2 ρs(t)] when the bond between site ` and `+1 The spectrum of the corresponding Lindbladian with a unique is steered.P In the following we will use the shorthand 0 0 zero-eigenvalue corresponding to the singlet state. (`,`+1) 0 (` ,` +1) (`,`+1) = Stot=2 , = Stot=2 , and Ui = Ui . P UsingP Eq. (B4P ), itP can be shown that

0 0 X † 0 2 Eq. (29), let us consider only one of them, such that Tr[ ρs(t + δt)] =Tr[ ρs(t)] + Tr [Un Unρ(t)] sin δt P P n P − 0 0 H = J(σ U + h.c.). (C1) Tr[( + )ρs(t)](1 cos δt) s-d d 3 − PP P P − ⊗ 0 2 + Tr[ ρs(t)](1 cos δt) . PP P − The above coupling steers the state of the system to the (D1) singlet only from the T0 state and leaves the T± sub- space untouched. However,| i the state can be shaken| iout of The above equation has three different cases the T± subspace onto the T0 via a system-Hamiltonian | i | i ` = `0: In this case Eq. (D1) reduces to of the form • 2 Tr[ ρs(t + δt)] = Tr[ ρs(t)] cos δt , (D2) Hs = θ1 σ1 + θ2 σ2, (C2) P P · · which shows that the bond (`, ` + 1) is steered to- where θ1/2 can be arbitrary vectors [100]. wards the AKLT state. The time evolution map of the density matrix of the ` and `0 are such that the two bonds do not share system can be written explicitly as • 0 a common site. In this case, Tr[ ρs(t + δt)] = 0 P Tr[ ρs(t)] since the steering has no effect on the   δt δt P 0 0 δt −iHs-d iHs-d reduced density matrix of the bond (` , ` + 1). ρs t + = Trd[e 2 ρd ρs(t)e 2 ], (C3a) 2 ⊗ ` and `0 are adjacent such that the two bonds in • question share a common site. In this case   0 2 −iH δt δt iH δt Tr[ (ρs(t + δt) ρs(t))] = 2 sin (δt/2)Tr[ ρs(t)] s 2 s 2 P − − A ρs (t + δt) = e ρs t + e , (C3b) 4 2 + 4 sin (δt/2)Tr[ ρs(t)], B (D3) with Hs-d and Hs given by Eqs. (C1) and (C2) respec- tively. Simulating the protocol described above again where shows steering to pure singlet state: see Fig.6(a). The 0 0 X † 0 = + 2 U Un (D4a) Lindblad equation describing the dynamics in the time- A PP P P − nP continuum limit now also has a unitary part in addi- n tion to the dissipative part. The spectrum of the corre- sponding Lindbladian has a unique zero eigenvalue: see X = 0 U † 0U . (D4b) Fig.6(b). n n B PP P − n P Such protocols can be easily generalised to the case of the AKLT state. For a pair of spins-1, a single system- detector coupling steers weight out of one of the Stot = 2 The change in energy of the system measured with states and a Hamiltonian rotates the system within the respect to HAKLT can be obtained by summing Eq. (D3) Stot = 2 subspace in a sufficiently general fashion. over `0 which gives 20

2 4 ∆EAKLT(t) = 2 sin (δt/2)Tr[(2 + − + +)ρs(t)] + 4 sin (δt/2)Tr[( + − + +)ρs(t)] , (D5) − P A A P B B

(`,`+1) 1.2 0.05 Tr[ Stot=2 ρs(t)] shows a non-monotonic behaviour. This is aP manifestation of the phenomenon that when the bond 1.0 0.00 adjacent to the one being monitored is steered, the for- mer is steered away from the target as such the total

0.8 0.05 energy grows. − )]

t 1.4 D

( 1 s ρ

0.6 ˆ DF AKLT 0.10 E [ E − 1.2 (3,4)

T r Tr[ tot ρ (t)] PS =2 s 0.4 0.15 − 1.0

0.2 0.20 0.8 − 0.6 0.0 0.25 − 100 101 102 103 steering events 0.4

0.2

FIG. 7. Evolution of Tr[Eˆρs(t)] (red,orange) and EAKLT (blue,cyan) for an initial state which is the eigenstate cor- 0.0 0 1 2 3 responding to the most negative eigenvalue of Eˆ. Results are 10 10 10 10 steering events shown for a N = 4 spin-1 chain. The dots correspond to data points after steering each bond whereas the squares cor- respond to data points after steering each bond of the system FIG. 8. Evolution of the global and local measures of dis- once. tance from the AKLT ground state for the initial state used in Fig.7 (the eigenstate corresponding to the most negative eigenvalue of Eˆ). The Frobenius distance from the target state where the operators ± and ± are defined in Eq. (D4) has a monotonic decay whereas the trace distance at early with `0 = ` 1. To leadingA orderB in δt, a sufficient condi- ± times does not change, although it does eventually decay to tion for the steering to be monotonic would be that four- zero. The reduced density matrix of a particular bond also site operator ˆ` = +( ++ −)/2 has only non-negative gets steered to that of the AKLT ground state but in a non- eigenvalues.E HoweverP itA turnsA out that the spectrum of monotonic fashion. Results are shown for a N = 4 spin-1 ˆ has both positive and negative eigenvalues. Denoting chain. Similar to Fig.7 the dots correspond to data points theE eigenvalues and eigenvectors of ˆ as w and w , the after steering each bond whereas the squares correspond to E | i data points after steering each bond of the system once. condition on ρs(t) for EAKLT(t) to decay (grow) is given ˆ P by Tr[ ρs(t)] = w w w ρs(t) w ≷ 0. ThisE observation allowsh | us| toi specifically construct density matrices which show non-monotonic steering as Turning to global measures of distance, the trace dis- measured by E . For example, consider the initial tance, D1(t), at early times is flat and it decays to zero AKLT only at late times. This is analogous to the situation de- density matrix as ρs(t = 0) = w− w− where w− is | i h | scribed in the main text where the overlap of the state the most negative eigenvalue of ˆ. For such a state with the AKLT ground state stays zero but the state Tr[ ˆρ (t)] is trivially negative. NumericalE results for such s keeps evolving. The Frobenius distance, D (t), on the a situationE are shown in Fig.7 where E (t) initially F AKLT other hand decays monotonically to zero. grows while Tr[ ˆρs(t)] < 0 and starts decaying only when E The above results show that for many-body systems, Tr[ ˆρs(t)] > 0. E monotonicity of steering and the exact condition for con- For such an initial state, it is also interesting com- vergence to the target state can depend on the dis- pare the behaviour of other distance measures, both spa- tance measure employed. Crucially, however, these re- (`,`+1) tially local such as Tr[ tot ρs(t)] and global such as sults along with the numerical results in Sec.IVB show PS =2 D1(t) and DF (t). The results are shown in Fig.8. The that the system is guaranteed to be steered to the AKLT distance of a local reduced density matrix measured via ground state.

[1] E. Schr¨odinger,“Discussion of probability relations be- Soc. 31, 555563 (1935). tween separated systems,” Math. Proc. Camb. Philos. 21

[2] E. Schr¨odinger, “Probability relations between sepa- (2011). rated systems,” Math. Proc. Camb. Philos. Soc. 32, [22] S. Diehl, E. Rico, M. A. Baranov, and P. Zoller, “Topol- 446452 (1936). ogy by dissipation in atomic quantum wires,” Nat. Phys. [3] E. Schr¨odinger, “Die erfassung der quantengesetze 7, 971 (2011). durch kontinuierliche funktionen,” Naturwissenschaften [23] F. Ticozzi and L. Viola, “Stabilizing entangled states 17, 486–489 (1929). with quasi-local quantum dynamical semigroups,” Phil. [4] H. M. Wiseman, S. J. Jones, and A. C. Doherty, Trans. R. Soc. A 370, 5259–5269 (2012). “Steering, Entanglement, Nonlocality, and the Einstein- [24] C.-E. Bardyn, M. A. Baranov, C. V. Kraus, E. Rico, Podolsky-Rosen paradox,” Phys. Rev. Lett. 98, 140402 A. Imamo˘glu,P.˙ Zoller, and S. Diehl, “Topology by (2007). dissipation,” New J. Phys. 15, 085001 (2013). [5] D. Cavalcanti and P. Skrzypczyk, “Quantum steering: a [25] Z. Leghtas, U. Vool, S. Shankar, M. Hatridge, S. M. review with focus on semidefinite programming,” Rep. Girvin, M. H. Devoret, and M. Mirrahimi, “Stabilizing Prog. Phys. 80, 024001 (2016). a bell state of two superconducting qubits by dissipation [6] R. Uola, A. Costa, H. C. Nguyen, and O. G¨uhne, engineering,” Phys. Rev. A 88, 023849 (2013). “Quantum steering,” arXiv preprint arXiv:1903.06663 [26] Y. Liu, S. Shankar, N. Ofek, M. Hatridge, A. Narla, (2019). K. M. Sliwa, L. Frunzio, R. J. Schoelkopf, and M. H. [7] J. Von Neumann, Mathematische grundlagen der quan- Devoret, “Comparing and combining measurement- tenmechanik (Springer-Verlag (Berlin), 1932). based and driven-dissipative entanglement stabiliza- [8] J. Von Neumann and R.T. Beyer (translator), Mathe- tion,” Phys. Rev. X 6, 011022 (2016). matical Foundations of Quantum Mechanics: New Edi- [27] G. Lindblad, “On the generators of quantum dynamical tion (Princeton university press, 2018). semigroups,” Comm. Math. Phys. 48, 119–130 (1976). [9] Y. Aharonov, D. Z. Albert, and L. Vaidman, “How the [28] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, result of a measurement of a component of the spin of “Completely positive dynamical semigroups of n-level a spin-1/2 particle can turn out to be 100,” Phys. Rev. systems,” J. Math. Phys. 17, 821–825 (1976). Lett. 60, 1351–1354 (1988). [29] C. Altafini and F. Ticozzi, “Modeling and control of [10] P. A. Mello and L. M. Johansen, “Measurements in quantum systems: An introduction,” IEEE Transac- quantum mechanics and von Neumanns model,” in AIP tions on Automatic Control 57, 1898–1917 (2012). Conference Proceedings, Vol. 1319 (AIP, 2010) pp. 3–13. [30] F. Ticozzi, L. Zuccato, P. D. Johnson, and L. Viola, [11] P. A. Mello, “The von Neumann model of measurement “Alternating projections methods for discrete-time sta- in quantum mechanics,” in AIP Conference Proceed- bilization of quantum states,” IEEE Transactions on ings, Vol. 1575 (AIP, 2014) pp. 136–165. Automatic Control 63, 819–826 (2017). [12] B. Tamir and E. Cohen, “Introduction to weak mea- [31] V. P. Belavkin, “Quantum stochastic calculus and quan- surements and weak values,” Quanta 2, 7–17 (2013). tum nonlinear filtering,” Journal of Multivariate analy- [13] B. E. Y. Svensson, “Pedagogical review of quantum sis 42, 171–201 (1992). measurement theory with an emphasis on weak mea- [32] Z. K. Minev, S. O. Mundhada, S. Shankar, P. Reinhold, surements,” Quanta 2, 18–49 (2013). R. Guti´errez-J´auregui,R. J. Schoelkopf, M. Mirrahimi, [14] T. A. Brun, “A simple model of quantum trajectories,” H. J. Carmichael, and M. H. Devoret, “To catch and Am. J. Phys. 70, 719–737 (2002). reverse a quantum jump mid-flight,” Nature 570, 200 [15] A. N. Korotkov, “Noisy quantum measurement of solid- (2019). state qubits: Bayesian approach,” in Quantum Noise in [33] I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, “Rig- Mesoscopic Physics (Springer, 2003) pp. 205–228. orous results on valence-bond ground states in antifer- [16] A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, romagnets,” Phys. Rev. Lett. 59, 799–802 (1987). and R. J. Schoelkopf, “Introduction to quantum noise, [34] I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, “Va- measurement, and amplification,” Rev. Mod. Phys. 82, lence bond ground states in isotropic quantum antifer- 1155–1208 (2010). romagnets,” Comm. Math. Phys. 115, 477–528 (1988). [17] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. [35] It is clear that a projective measurement of σz will not B¨uchler, and P. Zoller, “Quantum states and phases in guarantee steering to |↑i as the measurement would driven open quantum systems with cold atoms,” Nat. project the spin to either |↑i or |↓i with probabilities z Phys. 4, 878 (2008). Tr[ρs(t)(1 ± σ )/2] respectively, ρs(t) being the state of [18] B. Kraus, H. P. B¨uchler, S. Diehl, A. Kantian, the system spin-1/2. A. Micheli, and P. Zoller, “Preparation of entangled [36] F. D. M. Haldane, “Fractional quantization of the hall states by quantum markov processes,” Phys. Rev. A 78, effect: A hierarchy of incompressible quantum fluid 042307 (2008). states,” Phys. Rev. Lett. 51, 605–608 (1983). [19] M. Roncaglia, M. Rizzi, and J. I. Cirac, “Pfaffian state [37] A.Yu. Kitaev, “Fault-tolerant quantum computation by generation by strong three-body dissipation,” Phys. anyons,” Annals of Physics 303, 2 – 30 (2003). Rev. Lett. 104, 096803 (2010). [38] J. H. Son and J. Alicea, “Commuting-projector [20] H. Weimer, M. M¨uller,I. Lesanovsky, P. Zoller, and hamiltonians for chiral topological phases built from H. P. B¨uchler, “A Rydberg ,” Nat. parafermions,” Phys. Rev. B 97, 245144 (2018). Phys. 6, 382–388 (2010). [39] N. Tantivasadakarn and A. Vishwanath, “Full com- [21] B. P. Lanyon, C. Hempel, D. Nigg, M. M¨uller,R. Ger- muting projector hamiltonians of interacting symmetry- ritsma, F. Z¨ahringer, P. Schindler, J. T. Barreiro, protected topological phases of fermions,” Phys. Rev. B M. Rambach, G. Kirchmair, M. Hennrich, P. Zoller, 98, 165104 (2018)√ . R. Blatt, and C. F. Roos, “Universal digital quan- [40] Note that the δt in the rescaling of the energy-scales tum simulation with trapped ions,” Science 334, 57–61 22

associated to the operator U is strongly reminiscent of rial for high-impedance quantum circuits,” Nature Ma- the scaling of the white noise term in an Itˆostochastic terials 18, 816 (2019). differential equation [101]. This is not surprising as a [58] Nataliya Maleeva, Lukas Gr¨unhaupt, T Klein, F Levy- stochastic Schr¨odingerequation with white noise, when Bertrand, O Dupre, M Calvo, F Valenti, P Winkel, averaged over the noise, yields a Lindblad equation for F Friedrich, W Wernsdorfer, et al., “Circuit quantum the density matrix with the jump operators given by the electrodynamics of granular aluminum resonators,” Na- terms to which the noise couples [102]. ture communications 9 (2018). [41] Note that while the system now is composed of spins-1, [59] Abigail Shearrow, Gerwin Koolstra, Samuel J White- we continue to use spins-1/2 as detectors. ley, Nathan Earnest, Peter S Barry, F Joseph Heremans, [42] V. E. Korepin and Y. Xu, “Entanglement in valence- David D Awschalom, Erik Shirokoff, and David I Schus- bond-solid states,” Int. J. Mod. Phys. B 24, 1361 (2010). ter, “Atomic layer deposition of titanium nitride for [43] J. Dalibard, Y. Castin, and K. Mølmer, “Wave-function quantum circuits,” Applied Physics Letters 113, 212601 approach to dissipative processes in quantum optics,” (2018). Phys. Rev. Lett. 68, 580–583 (1992). [60] N. Akerman, S. Kotler, Y. Glickman, and R. Ozeri, [44] K. Mølmer, Y. Castin, and J. Dalibard, “Monte carlo “Reversal of photon-scattering errors in atomic qubits,” wave-function method in quantum optics,” J. Opt. Soc. Phys. Rev. Lett. 109, 103601 (2012). Am. B 10, 524 (1993). [61] Yinnon Glickman, Shlomi Kotler, Nitzan Akerman, [45] J. P. Groen, D. Rist`e,L. Tornberg, J. Cramer, P. C. and Roee Ozeri, “Emergence of a measurement ba- de Groot, T. Picot, G. Johansson, and L. DiCarlo, sis in atom-photon scattering,” Science 339, 1187–1191 “Partial-measurement backaction and nonclassical weak (2013). values in a superconducting circuit,” Phys. Rev. Lett. [62] Yiheng Lin, JP Gaebler, Florentin Reiter, Ting R Tan, 111, 090506 (2013). Ryan Bowler, AS Sørensen, Dietrich Leibfried, and [46] M. Hatridge, S. Shankar, M. Mirrahimi, F. Schackert, David J Wineland, “Dissipative production of a maxi- K. Geerlings, T. Brecht, K. M. Sliwa, B. Abdo, L. Frun- mally entangled steady state of two quantum bits,” Na- zio, S. M. Girvin, R. J. Schoelkopf, and M. H. De- ture 504, 415–418 (2013). voret, “Quantum back-action of an individual variable- [63] Daniel Kienzler, H-Y Lo, B Keitch, L De Clercq, F Le- strength measurement,” Science 339, 178–181 (2013). upold, F Lindenfelser, M Marinelli, V Negnevitsky, and [47] K. W. Murch, S. J. Weber, C. Macklin, and I. Siddiqi, JP Home, “Quantum harmonic oscillator state synthesis “Observing single quantum trajectories of a supercon- by reservoir engineering,” Science 347, 53–56 (2015). ducting quantum bit,” Nature 502, 211 (2013). [64] A. Retzker, private communication. [48] D. Tan, S. J. Weber, I. Siddiqi, K. Mølmer, and K. W. [65] J. Cai, F. Jelezko, M. B. Plenio, and A. Ret- Murch, “Prediction and retrodiction for a continuously zker, “Diamond-based single-molecule magnetic reso- monitored superconducting qubit,” Phys. Rev. Lett. nance spectroscopy,” New Journal of Physics 15, 013020 114, 090403 (2015). (2013). [49] Shay Hacohen-Gourgy, Leigh S. Martin, Emmanuel [66] Jianming Cai, Alex Retzker, Fedor Jelezko, and Mar- Flurin, Vinay V. Ramasesh, K. Birgitta Whaley, and Ir- tin B Plenio, “A large-scale quantum simulator on a fan Siddiqi, “Quantum dynamics of simultaneously mea- diamond surface at room temperature,” Nature Physics sured non-commuting observables,” Nature 538, 491– 9, 168–173 (2013). 494 (2016). [67] J. R. Maze, A. Gali, E. Togan, Y. Chu, A. Trifonov, [50] M. Naghiloo, M. Abbasi, Yogesh N. Joglekar, and E. Kaxiras, and M. D. Lukin, “Properties of nitrogen- K. W. Murch, “Quantum state tomography across the vacancy centers in diamond: the group theoretic ap- exceptional point in a single dissipative qubit,” Nature proach,” New Journal of Physics 13, 025025 (2011). Physics 15, 1232–1236 (2019). [68] A. Batalov, C. Zierl, T. Gaebel, P. Neumann, I.-Y. [51] O. Zilberberg, A. Romito, and Y. Gefen, “Null weak Chan, G. Balasubramanian, P. R. Hemmer, F. Jelezko, values in multi-level systems,” Physica Scripta 2012, and J. Wrachtrup, “Temporal coherence of photons 014014 (2012). emitted by single nitrogen-vacancy defect centers in di- [52] Oded Zilberberg, Alessandro Romito, David J. Star- amond using optical rabi-oscillations,” Phys. Rev. Lett. ling, Gregory A. Howland, Curtis J. Broadbent, John C. 100, 077401 (2008). Howell, and Yuval Gefen, “Null values and quantum [69] Tuvia Gefen, Maxim Khodas, Liam P. McGuinness, Fe- state discr`ımination,” Phys. Rev. Lett. 110, 170405 dor Jelezko, and Alex Retzker, “Quantum spectroscopy (2013). of single spins assisted by a classical clock,” Phys. Rev. [53] K. W. Murch, private communication. A 98, 013844 (2018). [54] I. M. Pop, private communication. [70] K. S. Cujia, J. M. Boss, K. Herb, J. Zopes, and C. L. [55] V. E. Manucharyan, J. Koch, L. I. Glazman, and M. H. Degen, “Tracking the precession of single nuclear spins Devoret, “Fluxonium: Single cooper-pair circuit free of by weak measurements,” Nature 571, 230–233 (2019). charge offsets,” Science 326, 113–116 (2009). [71] M. H. Abobeih, J. Randall, C. E. Bradley, H. P. [56] I. M. Pop, K. Geerlings, G. Catelani, R. J. Schoelkopf, Bartling, M. A. Bakker, M. J. Degen, M. Markham, L. I. Glazman, and M. H. Devoret, “Coherent suppres- D. J. Twitchen, and T. H. Taminiau, “Atomic-scale sion of electromagnetic dissipation due to superconduct- imaging of a 27-nuclear-spin cluster using a quantum ing quasiparticles,” Nature 508, 369 (2014). sensor,” Nature 576, 411–415 (2019). [57] L. Gr¨unhaupt,M. Spiecker, D. Gusenkova, N. Maleeva, [72] Massimiliano Rossi, David Mason, Junxin Chen, S. T. Skacel, I. Takmakov, F. Valenti, P. Winkel, Yeghishe Tsaturyan, and Albert Schliesser, H. Rotzinger, W. Wernsdorfer, A. V. Ustinov, and I. M. “Measurement-based quantum control of mechani- Pop, “Granular aluminium as a superconducting mate- cal motion,” Nature 563, 53–58 (2018). 23

[73] Shlomi Kotler, Gabriel A Peterson, Ezad Shojaee, Flo- tanglement transition from variable-strength weak mea- rent Lecocq, Katarina Cicak, Alex Kwiatkowski, Shawn surements,” Phys. Rev. B 100, 064204 (2019). Geller, Scott Glancy, Emanuel Knill, Raymond W Sim- [89] C.-M. Jian, Y.-Z/ You, R. Vasseur, and A. W. W. Lud- monds, et al., “Tomography of entangled macroscopic wig, “Measurement-induced criticality in random quan- mechanical objects,” arXiv preprint arXiv:2004.05515 tum circuits,” arXiv preprint arXiv:1908.08051 (2019). (2020). [90] M. J. Gullans and D. A. Huse, “Dynamical purification [74] R. Simon, “Peres-horodecki separability criterion for phase transition induced by quantum measurements,” continuous variable systems,” Phys. Rev. Lett. 84, arXiv preprint arXiv:1905.05195 (2019). 2726–2729 (2000). [91] S. Choi, X.-L. Bao, Y.and Qi, and E. Altman, “Quan- [75] Lu-Ming Duan, G. Giedke, J. I. Cirac, and P. Zoller, tum error correction and entanglement phase transi- “Inseparability criterion for continuous variable sys- tion in random unitary circuits with projective mea- tems,” Phys. Rev. Lett. 84, 2722–2725 (2000). surements,” arXiv preprint arXiv:1903.05124 (2019). [76] E. Karimi, private communication. [92] Y. Bao, S. Choi, and E. Altman, “Theory of the [77] Y. Wang, J. Li, S. Zhang, K. Su, Y. Zhou, K. Liao, phase transition in random unitary circuits with mea- S. Du, H. Yan, and S.-L. Zhu, “Efficient quantum mem- surements,” arXiv preprint arXiv:1908.04305 (2019). ory for single-photon polarization qubits,” Nature Pho- [93] Q. Tang and W. Zhu, “Measurement-induced phase tonics 13, 346–351 (2019). transition: A case study in the non-integrable model [78] Y.-W. Cho, Y. Kim, Y.-H. Choi, Y.-S. Kim, S.-W. Han, by density-matrix renormalization group calculations,” S.-Y. Lee, S. Moon, and Y.-H. Kim, “Emergence of arXiv preprint arXiv:1908.11253 (2019). the geometric phase from quantum measurement back- [94] F. Wilczek and A. Zee, “Appearance of gauge struc- action,” Nature Physics 15, 665–670 (2019). ture in simple dynamical systems,” Phys. Rev. Lett. 52, [79] Jing Zhang, Yu-xi Liu, Re-Bing Wu, Kurt Jacobs, and 2111–2114 (1984). Franco Nori, “Quantum feedback: theory, experiments, [95] K. Snizhko, R. Egger, and Y. Gefen, “Non-abelian geo- and applications,” Physics Reports 679, 1–60 (2017). metric dephasing,” Phys. Rev. Lett. 123, 060405 (2019). [80] N. Goldman, J. C. Budich, and P. Zoller, “Topologi- [96] Z. Gong, R. Hamazaki, and M. Ueda, “Discrete time- cal quantum matter with ultracold gases in optical lat- crystalline order in cavity and circuit qed systems,” tices,” Nature Physics 12, 639 (2016). Phys. Rev. Lett. 120, 040404 (2018). [81] F. Ciccarello, “Collision models in quantum optics,” [97] F. M. Gambetta, F. Carollo, M. Marcuzzi, J. P. Gar- Quantum Meas. Quantum Metrol. 4, 53 – 63 (2017). rahan, and I. Lesanovsky, “Discrete time crystals [82] V. Gebhart, K. Snizhko, T. Wellens, A. Buchleit- in the absence of manifest symmetries or disorder in ner, A. Romito, and Y. Gefen, “Topological transi- open quantum systems,” Phys. Rev. Lett. 122, 015701 tion in measurement-induced geometric phases,” arXiv (2019). preprint arXiv:1905.01147 (2019). [98] A. Lazarides, S. Roy, F. Piazza, and R. Moessner, “On [83] D. Gottesman, “Stabilizer codes and quantum error cor- time crystallinity in dissipative floquet systems,” arXiv rection,” arXiv preprint quant-ph/9705052 (1997). preprint arXiv:1904.04820 (2019). [84] Y. Li, X. Chen, and M. P. A. Fisher, “Quantum zeno ef- [99] J. O’Sullivan, O. Lunt, C. W. Zollitsch, M. L. W. The- fect and the many-body entanglement transition,” Phys. walt, J. J. L. Morton, and A. Pal, “Dissipative discrete Rev. B 98, 205136 (2018). time crystals,” arXiv preprint arXiv:1807.09884 (2018). [85] B. Skinner, J. Ruhman, and A. Nahum, “Measurement- [100] Exceptions are special choices of θ for which Hs has induced phase transitions in the dynamics of entangle- an eigenstate which lives purely in the |T±i subspace. ment,” Phys. Rev. X 9, 031009 (2019). Such a special choice does not transfer weight from that [86] A. Chan, R. M. Nandkishore, M. Pretko, and particular state in the |T±i to the |T0i and hence in G. Smith, “Unitary-projective entanglement dynamics,” the steady state, some weight remains in the orthogonal Phys. Rev. B 99, 224307 (2019). subspace. [87] Y. Li, X. Chen, and M. P. A. Fisher, “Measurement- [101] N. G. Van Kampen, Stochastic processes in physics and driven entanglement transition in hybrid quantum cir- chemistry, Vol. 1 (Elsevier, 1992). cuits,” arXiv preprint arXiv:1901.08092 (2019). [102] Z. Cai and T. Barthel, “Algebraic versus exponential de- [88] M. Szyniszewski, A. Romito, and H. Schomerus, “En- coherence in dissipative many-particle systems,” Phys. Rev. Lett. 111, 150403 (2013).