Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire...

14
Observing the space- and time-dependent growth of correlations in dynamically tuned synthetic Ising antiferromagnets Vincent Lienhard * , Sylvain de L´ es´ eleuc * , Daniel Barredo, Thierry Lahaye, and Antoine Browaeys Laboratoire Charles Fabry, Institut d’Optique Graduate School, CNRS, Universit´ e Paris-Saclay, F-91127 Palaiseau Cedex, France Michael Schuler * , Louis-Paul Henry, and Andreas M. L¨ auchli Institut f¨ ur Theoretische Physik, Universit¨ at Innsbruck, A-6020 Innsbruck, Austria (Dated: November 6, 2017) We explore the dynamics of artificial one- and two-dimensional Ising-like quantum antiferromagnets with different lattice geometries by using a Rydberg quantum simulator of up to 36 spins in which we dynamically tune the parameters of the Hamiltonian. We observe a region in parameter space with antiferromagnetic (AF) ordering, albeit with only finite-range correlations. We study systematically the influence of the ramp speeds on the correlations and their growth in time. We observe a delay in their build-up associated to the finite speed of propagation of correlations in a system with short-range interactions. We obtain a good agreement between experimental data and numerical simulations taking into account experimental imperfections measured at the single particle level. Finally, we develop an analytical model, based on a short-time expansion of the evolution operator, which captures the observed spatial structure of the correlations, and their build-up in time. I. INTRODUCTION The study of non-equilibrium dynamics is currently one of the most challenging areas of quantum many-body physics. In contrast to the equilibrium case, where statistical physics pro- vides a general theoretical framework and where very pow- erful numerical methods are available, the out-of-equilibrium behavior of quantum matter presents a wide variety of phe- nomena and is extremely hard to simulate numerically, es- pecially in dimensions d> 1. When the parameters of a quantum many-body system are quenched abruptly, or more generally ramped with a finite rate into a new quantum phase, correlations and entanglement build up and propagate over the system, which may, at long times, either thermalize or retain memory of its initial state when many-body localization oc- curs [1, 2]. The speed at which the correlations propagate depends on the range of the interaction and is limited by the Lieb-Robinson bounds [38]. An attractive way to study this physics has emerged in the last years, and consists in using quantum simulators, i.e. well- controlled, artificial quantum systems that implement exper- imentally the Hamiltonian of interest [9]. Spin Hamiltoni- ans that are used in condensed-matter physics to describe e.g. quantum magnets are arguably the simplest quantum many-body systems that can be used to study non-equilibrium dynamics: even though they involve distinguishable parti- cles with only internal degrees of freedom, the interplay be- tween interactions, geometry and dimensionality provides a wealth of distinct quantum phases into which the system can be driven. In recent years, many experimental platforms for the quantum simulation of spin Hamiltonians have been devel- oped using the tools of atomic physics. For example, equilib- rium properties of synthetic quantum magnets have been stud- ied using trapped ions [10, 11] or ultra-cold atoms in optical lattices [1216] including e.g. the observation of long-range antiferromagnetic order [17]. Many experiments using these platforms were also devoted to the study of non-equilibrium dynamics, including the investigation of the Lieb-Robinson bound [1824]. Recently, a new platform, using arrays of individually re- solved atoms excited to Rydberg states, has been shown to provide a versatile way to engineer synthetic quantum Ising magnets [25]. Pioneering experiments in quantum gas mi- croscopes studied quenches [26] or slow sweeps [27] in a regime where the blockade radius R b , i.e. the distance over which interatomic interactions prevent the excitation of two atoms, was much larger than the lattice spacing a, rendering the underlying lattice hardly relevant. In this case, the ob- served correlations are liquid-like, and observing the crystal- like ground state of the system [28] would require exponen- tially long ramps [29]. More recently, experiments with arrays of optical tweezers allowed exploring the regime R b & a, studying non-equilibrium dynamics following quenches [30] or slow sweeps [31]. Here, we use a Rydberg-based platform emulating an Ising antiferromagnet to study the growth of correlations during ramps of the experimental parameters in 1d and 2d arrays of up to 36 single atoms with different geometries. We op- erate in the regime R b a where the interactions act to a good approximation only between nearest neighbors. We dy- namically tune the parameters of the Hamiltonian and observe the build-up of antiferromagnetic order. We also observe the influence of the finite ramp speed on the extent of the cor- relations, and follow the development in space and time of these correlations during a ramp. Numerical simulations of the dynamics of the system without any adjustable parame- ters are in very good agreement with the experimental data, and show that single-particle dephasing arising from techni- cal imperfections currently limits the range of the observed correlations. Finally, we observe a characteristic spatial struc- ture in the correlations, which can be understood qualitatively by a short-time expansion of the evolution operator for both, square and triangular lattices. The paper is organized as follows. In Sec. II, we describe the experimental platform. After recalling the phase diagram arXiv:1711.01185v1 [quant-ph] 3 Nov 2017

Transcript of Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire...

Page 1: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

Observing the space- and time-dependent growth of correlationsin dynamically tuned synthetic Ising antiferromagnets

Vincent Lienhard∗, Sylvain de Leseleuc∗, Daniel Barredo, Thierry Lahaye, and Antoine BrowaeysLaboratoire Charles Fabry, Institut d’Optique Graduate School, CNRS,

Universite Paris-Saclay, F-91127 Palaiseau Cedex, France

Michael Schuler∗, Louis-Paul Henry, and Andreas M. LauchliInstitut fur Theoretische Physik, Universitat Innsbruck, A-6020 Innsbruck, Austria

(Dated: November 6, 2017)

We explore the dynamics of artificial one- and two-dimensional Ising-like quantum antiferromagnets withdifferent lattice geometries by using a Rydberg quantum simulator of up to 36 spins in which we dynamicallytune the parameters of the Hamiltonian. We observe a region in parameter space with antiferromagnetic (AF)ordering, albeit with only finite-range correlations. We study systematically the influence of the ramp speedson the correlations and their growth in time. We observe a delay in their build-up associated to the finite speedof propagation of correlations in a system with short-range interactions. We obtain a good agreement betweenexperimental data and numerical simulations taking into account experimental imperfections measured at thesingle particle level. Finally, we develop an analytical model, based on a short-time expansion of the evolutionoperator, which captures the observed spatial structure of the correlations, and their build-up in time.

I. INTRODUCTION

The study of non-equilibrium dynamics is currently one ofthe most challenging areas of quantum many-body physics. Incontrast to the equilibrium case, where statistical physics pro-vides a general theoretical framework and where very pow-erful numerical methods are available, the out-of-equilibriumbehavior of quantum matter presents a wide variety of phe-nomena and is extremely hard to simulate numerically, es-pecially in dimensions d > 1. When the parameters of aquantum many-body system are quenched abruptly, or moregenerally ramped with a finite rate into a new quantum phase,correlations and entanglement build up and propagate over thesystem, which may, at long times, either thermalize or retainmemory of its initial state when many-body localization oc-curs [1, 2]. The speed at which the correlations propagatedepends on the range of the interaction and is limited by theLieb-Robinson bounds [3–8].

An attractive way to study this physics has emerged in thelast years, and consists in using quantum simulators, i.e. well-controlled, artificial quantum systems that implement exper-imentally the Hamiltonian of interest [9]. Spin Hamiltoni-ans that are used in condensed-matter physics to describee.g. quantum magnets are arguably the simplest quantummany-body systems that can be used to study non-equilibriumdynamics: even though they involve distinguishable parti-cles with only internal degrees of freedom, the interplay be-tween interactions, geometry and dimensionality provides awealth of distinct quantum phases into which the system canbe driven. In recent years, many experimental platforms forthe quantum simulation of spin Hamiltonians have been devel-oped using the tools of atomic physics. For example, equilib-rium properties of synthetic quantum magnets have been stud-ied using trapped ions [10, 11] or ultra-cold atoms in opticallattices [12–16] including e.g. the observation of long-rangeantiferromagnetic order [17]. Many experiments using theseplatforms were also devoted to the study of non-equilibrium

dynamics, including the investigation of the Lieb-Robinsonbound [18–24].

Recently, a new platform, using arrays of individually re-solved atoms excited to Rydberg states, has been shown toprovide a versatile way to engineer synthetic quantum Isingmagnets [25]. Pioneering experiments in quantum gas mi-croscopes studied quenches [26] or slow sweeps [27] in aregime where the blockade radius Rb, i.e. the distance overwhich interatomic interactions prevent the excitation of twoatoms, was much larger than the lattice spacing a, renderingthe underlying lattice hardly relevant. In this case, the ob-served correlations are liquid-like, and observing the crystal-like ground state of the system [28] would require exponen-tially long ramps [29]. More recently, experiments with arraysof optical tweezers allowed exploring the regime Rb & a,studying non-equilibrium dynamics following quenches [30]or slow sweeps [31].

Here, we use a Rydberg-based platform emulating an Isingantiferromagnet to study the growth of correlations duringramps of the experimental parameters in 1d and 2d arraysof up to 36 single atoms with different geometries. We op-erate in the regime Rb ' a where the interactions act to agood approximation only between nearest neighbors. We dy-namically tune the parameters of the Hamiltonian and observethe build-up of antiferromagnetic order. We also observe theinfluence of the finite ramp speed on the extent of the cor-relations, and follow the development in space and time ofthese correlations during a ramp. Numerical simulations ofthe dynamics of the system without any adjustable parame-ters are in very good agreement with the experimental data,and show that single-particle dephasing arising from techni-cal imperfections currently limits the range of the observedcorrelations. Finally, we observe a characteristic spatial struc-ture in the correlations, which can be understood qualitativelyby a short-time expansion of the evolution operator for both,square and triangular lattices.

The paper is organized as follows. In Sec. II, we describethe experimental platform. After recalling the phase diagram

arX

iv:1

711.

0118

5v1

[qu

ant-

ph]

3 N

ov 2

017

Page 2: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

2

FIG. 1. Studying the AF Ising model on 1d and 2d systems. (a) Examples of single-shot fluorescence images of single-atom arrays used inour experiments: a 24-atom 1d chain with periodic boundary conditions, a 6 × 6 square array, and a 36-atom triangular array. Each atomis used to encode a spin-1/2, whose internal states |↑〉 and |↓〉 are coupled with Rabi frequency Ω and detuning δ. (b) Time dependence ofthe Rabi frequency Ω(t) and detuning δ(t) used to probe the build-up of correlations. (c) Ground state phase diagrams of the Ising modelEq. (1), in the nearest-neighbor interaction limit, for a 1d chain, a 2d square lattice, and a 2d triangular lattice. AFM: antiferromagnetic. PM:paramagnetic. OBD: “order by disorder”. (d) Typical experimental correlation functions obtained for these geometries (see text). For the 1dchain the correlation length ξ = 1.5 sites (bottom left panel).

for the different array geometries (Sec. III), we explore it forthe square array (Sec. IV). In Sec. V A we study the influenceof ramp speeds on the correlations. In Sec. V B, we observea delay in the build-up of the spin-spin correlations, a featurelinked to their finite speed of propagation. Finally, in Sec. V C,we analyze the 2d spatial structure of the AF correlations onthe square and triangular geometries and show that it is qual-itatively captured by an analytical model based on short-timeexpansion.

II. EXPERIMENTAL PLATFORM

Our experimental platform (see Appendix A) is based onuser-defined two-dimensional arrays of optical tweezers, eachcontaining a single 87Rb atom [30]. Here we use the arraysshown in Fig. 1(a) containing up to N = 36 atoms: a 1dchain with periodic boundary conditions (PBC), a square lat-tice, and a triangular lattice. We achieve full loading of the

arrays using our atom-by-atom assembler [32]. The atoms areprepared in the ground state |↓〉 =

∣∣5S1/2, F = 2,mF = 2⟩

by optical pumping, and then coupled coherently to the Ryd-berg state |↑〉 =

∣∣64D3/2,mj = 3/2⟩

with a two-photon tran-sition of Rabi frequency Ω and a detuning δ, while the trapsare switched off. The system is described by the Hamiltonian:

H =∑i

(~Ω(t)

2σxi − ~δ(t)ni

)+

1

2

∑i 6=j

Uijninj , (1)

where ni = |↑〉 〈↑|i is the projector on the Rydberg state foratom i, and σx = |↑〉 〈↓| + |↓〉 〈↑| is the x-Pauli matrix. Theinteraction termUij arises from van der Waals interactions be-tween the atoms, and thus scales as 1/r6

ij with the distance rijbetween atoms i and j. This short-range character allows usto neglect interactions beyond nearest-neighbor (NN) atomsfor this work [33], and we thus restrict Eq. (1) to NN termsonly. For the D states we use, the van der Waals interac-tion is anisotropic [25, 34], and the lattice spacings in thearrays are tuned such that the NN interactions anisotropy is

Page 3: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

3

below 10%. We use typical values Ωmax/(2π) ∼ 2 MHz andU/h ∼ 1−3 MHz, see Table I. The driving parameters Ω andδ can be considered being constant over the entire array (seeAppendix A).

The system thus realizes an Ising-like model with a trans-verse field ∝ Ω and a longitudinal field ∝ δ, which givesrise to AF order for U > 0 (see below). We probe the sys-tem by using time-dependent ramps Ω(t) and δ(t) as shownin Fig. 1(b). The Rabi frequency Ω is switched on and offin trise, tfall ∼ 250 ns at a constant detuning δ, and, in be-tween, δ is ramped linearly from δ0/(2π) = −6 MHz to δfinal

during the time tsweep. The total duration of the ramp is thenttot = trise + tsweep + tfall. After this, the trap array is switchedon again. Atoms in |↓〉 are observed by fluorescence, whilethose in |↑〉 are lost from their trap. For a given set of pa-rameters, the experiment is repeated a few hundred times toreconstruct quantities of interest such as the Rydberg density(equivalent to the magnetization), spin-spin correlation func-tions, or the sublattice density histogram. When δfinal lies inthe AF region (see Fig. 1c), correlation functions show theemergence of short-range order [Fig. 1(d)] (see more detailson the measurements in Sec IV).

III. THEORETICAL PHASE DIAGRAMS AND STATEPREPARATION CONSIDERATIONS

The calculated ground state phase diagrams for the nearest-neighbor Ising model are shown in Fig. 1(c) for the three ge-ometries considered in this paper. On the 1d chain, the phasediagram is well-known and features an AF phase and a para-magnetic phase delimited by a second-order quantum phasetransition line of the (1+1)d Ising universality class [35]. For(~δ/U)TFI = z/2, with the coordination number z = 2, theHamiltonian corresponds exactly to the analytically solvabletransverse field Ising (TFI) model without longitudinal field,where in 1d the critical point (~Ω/U)c = 1/2 is known ana-lytically.

The phase diagram for the NN Hamiltonian on the squarelattice (z = 4) is qualitatively similar to the phase diagramon the chain, with again an AF phase and a paramagneticphase delimited by a second-order quantum phase transitionline in the (2+1)d Ising universality class. For the trans-verse field Ising line (~δ/U)TFI, the critical point is knownto high precision from Monte-Carlo simulations, (~Ω/U)c =1.52219(1) [36]. Finally, on the triangular lattice (z = 6) theNN Hamiltonian features a much richer phase diagram. A Ry-dberg crystal with filling fraction 1/3 (at Ω = 0), where one ofthe three triangular sublattices is occupied by Rydberg statesand the atoms in the other sublattices remain in the groundstate, appears in a region within 0 < ~δ/U < z/2. The conju-gate crystals obtained by flipping all the spins (filling fraction2/3 at Ω = 0) lies in the region within z/2 < ~δ/U < z. Themost interesting feature occurs when ~δ/U = z/2, where an“order by disorder” process occurs (see Appendix B). Whenincluding the open boundary conditions for the square and tri-angular N = 36 arrays studied here, the “phase diagrams”present the same phases with some modifications (see Ap-

pendix C).In the experiments reported here, we initialize the system in

the product state with all atoms in their ground state |↓〉. Sincethe system is not at equilibrium with a thermal bath, we cannotsimply cool to the ground state. Therefore we use an exper-imental protocol involving sweeps of Ω(t) and δ(t) in orderto characterize the ground states. Ideally, we would use adi-abatic state preparation to reach the targeted ground state. Inorder for this approach to succeed, the duration of the sweep,which scales as the inverse of the square of the energy gap ∆above the instantaneous ground states [37], should be smallerthan the coherence time of the system. In the 1d chain andthe 2d square lattice, the minimal gap at the quantum phasetransition scales as the inverse of the linear size of the system:∆ ∼ 1/N in 1d and ∆ ∼ 1/L = 1/

√N in 2d [35, 38]. The

gap also depends on the excitation velocity at the quantumcritical point, which can vary substantially along the quantumphase transition lines. For the triangular lattice, we expect afirst-order quantum phase transition between the paramagnetto either the 1/3 or 2/3 filling Rydberg crystals, resulting in aminimal gap exponentially small inN [39]. Due to these scal-ings, the gaps are small for the number of atoms used here. Asa consequence adiabatic state preparation would require longpulses. In the absence of any imperfections such as dephas-ing, pulse durations ttot of a few µs would allow to reachstrong correlations extending over the entire system. Whilesuch durations are experimentally accessible, state-of-the-artplatforms [27, 30, 31] show significant dephasing over thesetimescales. For this reason, we approach here the question ofstate preparation from the opposite side: we ask how muchcorrelations and what kind of structures we can expect beingbuilt up in a given amount of time. Answering these ques-tions also informs us about the minimal time required to buildup highly correlated states starting from a product state [4],i.e. about a quantum speed limit in our many-body system.

IV. EXPLORING THE SQUARE LATTICE PHASEDIAGRAM

In a first set of experiments we map out the Ω = 0 “phasediagram” on a L × L square array with L = 4 and L = 6.In order to do so, we use the ramps shown in Fig. 1(b) with~Ωmax/U = 2.3. This value is larger than the critical point(~Ω/U)c ≈ 1.5, such that the quantum phase transition lineis crossed while ramping down Ω, see Fig. 1(c). From theanalysis of the final fluorescence images, we reconstruct theRydberg density n =

∑i 〈ni〉 /N , and the connected spin-

spin correlation function

g(2)(k, l) =1

Nk,l

∑(i,j)

[〈ninj〉 − 〈ni〉 〈nj〉] , (2)

where the sum runs over atom pairs (i, j) whose separationis ri − rj = (ka, la), and Nk,l is the number of such atompairs in the array [40]. For a perfect antiferromagnetic Neelordering, g(2) takes the values±1/4 for |k|+ |l| even and odd,respectively.

Page 4: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

4

FIG. 2. Exploring the Ω = 0 line of the phase diagram for a L×L square lattice with L = 4 and L = 6. (a) Experimental spin-spin correlationfunction g(2)(k, l) for different values of the final detuning. (b) Corresponding experimental histograms of the populations of the two Neelsublattices A and B, to be contrasted with those in (c), calculated for uncorrelated random states with the same Rydberg density. (d) AverageRydberg density n, and (e) Neel structure factor SNeel as a function of the final detuning. The shaded areas highlight the region of the AF phasein the phase diagram for Ω = 0.

Figure 2(a) shows the spin-spin correlation function at theend of the ramp as a function of x = ~δfinal/U . We observestrong AF correlations, i.e. the sign of g(2)(k, l) changes ac-cording to the parity of the Manhattan distance |k| + |l|, inthe region 0 < x < 4 where AF order is expected, whilethe correlations vanish outside of this region. The ampli-tude of the AF correlations decreases with distance, in a waywhich is well captured by an exponential decay with a correla-tion length ξ, defined by g(2)(k, l) ∝ (−1)|k|+|l| exp[−(|k|+|l|)/ξ], of about 1.5 sites [see Figs. 1(d) and 3(b)]. Repeatingthe same experiment with a 1d chain yields the same corre-lation length (the particularity of the triangular lattice is dis-cussed in Appendix G). Importantly, even though the correla-tion length is smaller than two sites for both the 1d chain andthe square lattice, we are able to detect finite correlations withthe expected sign structure for up to five Manhattan shells, i.e.,almost over the whole array.

Another way to highlight AF correlations is to partition thearray into the two Neel sublattices A and B and plot a two-dimensional histogram P (nA, nB) with the |↑〉 populationsnA and nB of each sublattice as axes. For a perfect AF or-dering, one would observe populations only in the two cor-ners (nA = 0, nB = N/2) and (nA = N/2, nB = 0). Theexperimental results in Fig. 2(b) show that for x = 2.5 (cen-tral plot), the sublattice population histogram is substantiallyelongated along the anti-diagonal nB = N/2 − nA, which isnot observed for x < 0 or x > 4. For comparison, Fig. 2(c)shows the corresponding histograms that would be obtained

for an uncorrelated random state with the same average Ry-dberg density: there the elongation along the anti-diagonal isabsent.

In Fig. 2(d,e) we locate the boundaries of the AF phase.Panel (d) shows the mean Rydberg density n. For the Ω = 0ground state of Eq. (1) in the NN limit with periodic boundaryconditions, it should rise in steps, from 0 for x < 0, to 1/2 for0 < x < 4, and to 1 for x > 4 [41]. For open boundaryconditions, additional steps are present in the 0 < x < 4region, see Appendix C. Experimentally, the curve n(x) variescontinuously due to the finite duration of the sweep, as alsoobserved in Refs. [20, 27]. This mean density is therefore nota good observable to differentiate the paramagnetic and AFregions. Instead, we introduce the Neel structure factor SNeelto detect antiferromagnetic correlations in the AF region ofthe phase diagram:

SNeel = 4×∑k,l

(−1)|k|+|l|g(2)(k, l) . (3)

This quantity is an estimator for the correlation volume, i.e.,the number of spins correlated antiferromagnetically with agiven spin. In a situation with true long range order, SNeel di-verges linearly with the total “volume”N of the system, whileit stays almost constant for short-range ordered correlationswhen the system sizes are larger than the correlation length.As shown in Fig. 2(e), this quantity indicates the presence ofsubstantial short-range AF correlations for 0 < x < 4, asexpected from the phase diagram.

Page 5: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

5

FIG. 3. Build-up and decay of correlations in a 2d square lattice antiferromagnet for different ramp durations ttot. (a) Experimental resultsfor the 2d correlation function g(2)(k, l) for four different values of ttot, showing short-range AF ordering, with a pronounced sweep durationdependence. (b) The correlation function g(2) (averaged within a Manhattan shell m = |k| + |l|) decays exponentially with the Manhattandistance m, with a correlation length ξ of about 1.4 lattice sites. (c) Neel structure factor SNeel as a function of ttot. The green solid lineshows the result of a numerical simulation using exact diagonalization for a 4 × 4 system without dephasing, while the yellow dashed linedenotes a numerical simulation including dephasing (see Appendix D). (d) Dependence of the correlations on ttot for first-, second-, third- andfourth-neighbors. One observes a dependence not only on the Manhattan distance |k|+ |l|, but also on k − l (see Sec. V C). The dashed linesis a numerical simulation including dephasing (the shaded area corresponds to the s.e.m. obtained therein).

The results presented so far demonstrate that by analyzingthe correlation functions we can locate phase boundaries inthe phase diagrams. In the next section, we will study how thecorrelations build up in time.

V. EXPLORING THE TIME- AND SPACE-DEPENDENCEOF CORRELATIONS

In the following we first investigate in Sec. V A how SNeeldepends on the ramp speed and find a duration ttot that max-imizes its value. Then, with these settings, we study inSec. V B the temporal build-up of correlations during the op-timized sweep and observe delays between the growth of cor-relations at increasing distances. Eventually, in Sec. V C, weanalyze the spatial structure of these correlations and give aqualitative explanation of the observed patterns in terms of ashort-time expansion.

A. Correlations after ramps of varying durations

We perform experiments on a 36-site square lattice us-ing the sweep shown in Fig. 1(b) with fixed parameters~δfinal/U = 1.7, ~Ωmax/U = 0.7 < (~Ω/U)c. In contrast tothe previous experiment, the quantum phase transition line iscrossed while ramping the detuning δ and we vary the durationof the sweep tsweep (and therefore ttot). The initial parametersare chosen such that the ground state at t = 0 corresponds to|↓↓ · · · 〉 (no Rydberg excitation) and the ground state at thefinal parameters is an AF [42].

Fig. 3(a) shows two-dimensional plots of the experimentalg(2) correlation functions for four different values of ttot. Forthe shortest sweeps, the correlations remain weak. For inter-mediate durations (ttot ∼ 1µs) strong correlations emerge,with a staggered structure extending over many Manhattanshells, as shown in Fig. 3(b). For even larger sweep dura-tions, the correlation signal decreases with increasing sweepduration. In Fig. 3(c) we show the value of SNeel as a functionof ttot: starting from small values for short ramps, it exhibitsa broad maximum for ttot ∼ 1µs and slowly decreases forlonger ramp durations.

In order to gain a better understanding of the behavior de-

Page 6: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

6

0.5 0.6 0.7 0.8 0.9 1.0

t (µs)

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.354×

(−1)|k|+|l|g

(2) (k,l

)(k, l) = (0, 1)

(k, l) = (0, 2)

(k, l) = (1, 1)

(k, l) = (0, 3)

(k, l) = (2, 1)

(k, l) = (2, 2)

(k, l) = (3, 1)

(k, l) = (0, 4)

0.625 0.65 0.675 0.7 0.725 0.75 0.775 0.8

t (µs)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

nor

mal

ized m

=1

m=

2

m=

3

0

5

10

SN

eel

Experiment

Sim.

Sim. dephasing

0.4 0.6 0.8 1.0

t (µs)

0.0

0.5

1.0

1.5

2.0

ξ

(a) (b) (c)

(d)

FIG. 4. Time evolution of a 6 × 6 square array along the optimized ramp (see text). (a) Build-up of the correlation along the ramp. (b)Observation of a time delay between the build-up of significant correlations in increasing Manhattan shells m. This is a manifestation of aLieb-Robinson type limitation in the build-up of correlations between distant sites (see text). (c) Neel structure factor and, (d) correlationlength ξ. All figures: the dotted (dashed) lines correspond to the result of a numerical exact diagonalization simulation for a 4 × 4 latticewithout (with) dephasing.

scribed above, we compare the data with numerical simula-tions of the dynamics of the system governed by the Hamilto-nian of Eq. (1) with the full van der Waals interactions for thesweeps used in the experiment [43]. We show in Fig. 3(c) theresults of the simulation on a 4 × 4 square lattice. The greenline corresponds to a unitary evolution which models well theexperiment only for short sweep durations. The dashed yel-low line includes the decoherence observed in the experimenton the excitation of one atom through an empirical dephasingmodel (see Appendix D 2). We observe a remarkable agree-ment between this local (i.e. single atom) dephasing modelwith no adjustable parameters and the experiment over a largerange of ttot. This indicates that the saturation and the decayof the correlations for longer sweeps is due to (single-particle)decoherence.

We now analyze the spin-spin correlations for differentManhattan distances |k| + |l|. In Fig. 3(d) we observe thebuild up of correlations up to the fourth shell |k|+ |l| = 4, allof them being antiferromagnetically staggered, g(2)(k, l) ∝(−1)|k|+|l|. For short sweeps the correlations sharply increasewith increasing duration, saturate for longer sweeps, and de-cay, again due to decoherence. The simulation including de-phasing (dashed lines) reproduces well this trend.

B. Build-up of correlations along a ramp of fixed duration

In a subsequent experiment, we analyze how the AF corre-lations build up in time, during the sweep that maximizes SNeel(ttot = 1.0µs) identified in the previous subsection. Stoppingthe dynamics after a variable time 0 < t < ttot, Fig. 4 showsthe time evolution of the correlations after entering the AF re-gion identified in Fig. 2 at δ(t) > 0 (t > 0.5µs). We observethat most of the correlations build up for 0.6 < t < 0.8µsand then freeze at larger times. When comparing again to thenumerical simulation of the dynamics including the single-particle dephasing, we obtain a remarkable agreement withthe data.

Fig. 4 also features a striking effect: we observe a delayin the build-up of correlations between the different Manhat-tan shells m = |k| + |l|, highlighting the finite speed for thespread of correlations. To quantify this effect we normalizethe correlations for each distance such that the correspondingdephasing-free simulation reaches a maximum of 1 at largetime [Fig. 4(b)]. Fixing an arbitrary threshold of 0.2, we ob-serve that the nearest neighbor shell (m = 1) reaches thisthreshold at t ≈ 0.64µs, the second shell (m = 2) at t ≈0.71µs and finally the third shell (m = 3) at t ≈ 0.79µs (seegray vertical lines). This delay is a manifestation of the finitepropagation speed of correlations as theoretically predicted byLieb-Robinson bounds [3, 4]. Such bounds are well exploredin the context of quench dynamics [18, 20, 23, 44, 45], buttheir importance for state preparation protocols is not equallywell known (see Appendix E for a brief review on Lieb-Robinson bounds adapted for the present context).

Fig. 4(b) also reveals that when the correlations on thefirst shell reach the threshold, correlations for higher Manhat-tan distances are suppressed, but still detectable for m = 2.This illustrates the known fact that the correlations outsidethe Lieb-Robinson cone are finite, albeit exponentially sup-pressed (see also Appendix E). To recover this fact, we in-troduce a simple analytical approach based on a short-timeexpansion method (see Appendix F for details and specificresults). We find analytical expressions in powers of the dura-tion T of the ramp for the connected g(2) correlation func-tions for different Manhattan distance m valid in the limit(UT/~,ΩT, δT 1). We obtain nearest-neighbor correla-tions (m = 1) of order T 6 and next-nearest neighbor corre-lations (m = 2) of order T 10. More generally, the leadingorder of the expansion for two sites separated by a Manhattandistance m appears at order T 2+4m thus suggesting that at agiven (short) time the correlation decreases exponentially withm (see more details in Appendix F). This scaling explainsqualitatively the observed time dependence in Figs. 4(a) and(b), even though strictly speaking the range of applicabilityof the short-time expansion is limited to times much shorter

Page 7: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

7

FIG. 5. Spatial structure of the spin-spin correlations on the 36-site square (a-b) and on the 36-site triangular lattice (c-d). Experimental dataafter a sweep is shown in (a) and (c) together with the values of (k, l) for the first three shells |k| + |l| (see Fig. 1(d) for the full range). Thenumbers in brackets [c] give the number of linking paths contributing to the short-time expansion. (b) and (d) dependence of the correlation on(k, l) for several Manhattan distances with a qualitative comparison to binomial coefficients (green dotted lines) and leading-order short-timeexpansion (red dashed lines), see text for details. The shaded region corresponds to the s.e.m. of the experimental data.

than those shown there (for which the correlations would stillbe extremely small, and thus very challenging to measure ex-perimentally). However, our numerical exact diagonalizationresults without and with dissipation as well as the experimentshow that the qualitative features of the short-time expansionprevail for the considered non-perturbative times and the ac-tual ramp shapes.

C. Build-up of spatial structures on the square and triangularlattices

We finally analyze the spatial structure of the correlationsin more detail. Both Fig. 3(d) and Fig. 4(a,b) show that thecorrelations do not depend only on the Manhattan distancem = |k|+ |l|, but also, for fixed m, on k− l. For instance, forsecond neighbors (m = 2), the correlations for (k, l) = (0, 2)or (2, 0) is about half of those along the diagonal (k, l) =(1, 1). This spatial structure, absent in a 1d setting, has not, toour knowledge, been experimentally observed in 2d.

The observed spatial structures in the correlations within agiven Manhattan shell m can also be captured by the short-time expansion. The leading order coefficient of a given cor-relator g(2)(k, l) depends on the number of paths on the latticelinking the two sites as detailed in Appendix F 2. Fig. 5 showsa comparison of the spatial structure in experimental data forboth a square and a triangular lattice after a ramp of finite du-ration. In panels (a) and (c), the number of linking paths is

given for k ≥ l ≥ 0 by the binomial coefficient Cmk−l for both

lattices and is shown in brackets. In panels (b) and (d) weanalyze the correlations on the square and triangular lattice.The green dotted lines show the k− l dependence of g(2)(k, l)assumed to be given only by the binomial coefficients, nor-malized to the maximal correlation value for each subplot.

The precision of the experiment even allows us to observea small asymmetry in each shell m = const., due to a slightresidual anisotropy of the interaction. An extension of theshort-time expansion to the anisotropic case yields an ana-lytical expression from which we extract the ratio Uz/Uw

of the nearest-neighbor interactions. We use this interactionanisotropy in the analytical expression to obtain the correla-tion for larger |k| + |l|. The results are shown as red dashedlines.

Interestingly, for triangular lattices, in contrast with the caseof square arrays, the spatial structures of the correlations thatone observes experimentally, and which are well reproducedby the short-time expansion, do not reflect directly those onewould obtain in the ground state (see Appendix G). A detailedstudy of this specific behavior, which may be a dynamical sig-nature of the frustrated character of the triangular geometry, ishowever beyond the scope of the present work.

Page 8: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

8

VI. CONCLUSION AND OUTLOOK

In conclusion, we have used a Rydberg-based platform tostudy antiferromagnetic correlations in a synthetic Ising mag-net with different geometries. Using dynamical variations ofthe parameters, we explored the phase diagram and preparedarrays exhibiting antiferromagnetic order with pronouncedManhattan structures. We also studied the growth of the cor-relations during the sweeps of the parameters. We observeddelays in the build-up of correlations between sites at differ-ent distances, a feature linked to the Lieb-Robinson boundsfor the propagation of correlations in a system with nearest-neighbor interactions. We were able to understand the spatialstructure of the correlations after short to intermediate evo-lution times using an analytical short-time expansion of theevolution operator. Finally, we obtained remarkable agree-ment between the data on the dynamics of the correlationsand numerical simulations using a local (i.e. single particle)dephasing model.

In the future, the time over which coherent simulation canbe performed in such artificial quantum magnets could beextended by a better understanding of the dephasing mech-anisms at play in the coherent excitation of Rydberg states[46]. We will then be able to explore geometric frustrationon triangular or Kagome lattices [47, 48]. Another promisingavenue is the study of magnets described by the XY model im-plemented using resonant dipole-dipole interactions [49, 50]:there the coherent drive is obtained by using microwaves, andmuch lower dephasing rates are expected; moreover the inter-action decays slowly, as 1/r3, and long-range effects shouldbe more prominent [51].

Note: Similar antiferromagnetic correlations in 2d havebeen observed recently at Princeton University, using a sys-tem of Li Rydberg atoms in a quantum gas microscope.

ACKNOWLEDGMENTS

This work benefited from financial support by theEU [H2020 FET-PROACT Project RySQ], by the “In-vestissements d’Avenir” LabEx PALM (ANR-10-LABX-0039-PALM, projects QUANTICA and XYLOS) and by theRegion Ile-de-France in the framework of DIM Nano-K. MS,LPH and AML acknowledge support by the Austrian ScienceFund for project SFB FoQus (F-4018). The computationalresults presented have been achieved in part using the Vi-enna Scientific Cluster (VSC). This work was supported bythe Austrian Ministry of Science BMWF as part of the UniIn-frastrukturprogramm of the Focal Point Scientific Computingat the University of Innsbruck.∗ V. L., S. de L. and M. S. contributed equally to this work.

Appendix A: Experimental details

Here we give more details about our experimental setup andon the mapping of the Rydberg system onto an Ising antifer-romagnet.

FIG. 6. Influence of the sign of C6 on the many-body spectrum.Eigenenergies of a small 2 × 2 square system as a function of δ for~Ω/U ' 0.5. (a) For a repulsive interaction C6 > 0, the groundstate of H is antiferromagnetic for 0 < ~δ < 2U . It can be reachedfrom |↓〉⊗4 by adiabatically following the ground state starting fromnegative values of δ (trajectory in dashed arrows). (b) For an at-tractive interaction C6 < 0, the antiferromagnetic state is the mostexcited state, and the sign of the detuning needed to reach it from|↓〉⊗4 is reversed.

We use a two-photon excitation scheme to excite the atomsto the Rydberg state. The two lasers at 795 nm and 475 nm,with corresponding Rabi frequencies Ωr and Ωb, and a detun-ing ∆ ' 2π×740 MHz from the intermediate

∣∣5P1/2, F = 2⟩

state result in an effective Rabi frequency Ω = ΩrΩb/(2∆)for the coupling between |↓〉 and |↑〉, but also introduce light-shifts (Ω2

r−Ω2b)/(4∆) that add up to the two-photon detuning

δ. In order to generate the ramps shown in Fig. 1(b), we thuscompensate for these additional lightshifts. An AOM is usedfor changing dynamically the amplitude and frequency of thered beam. Due to the finite size of our excitation beams, theatoms do not all experience exactly the same δ and Ω. Forour arrays with a size ∼ 40 µm, the inhomogeneities of Ω arebelow 15%, while the detuning does not differ from its valueon the central atom by more than 150 kHz.

The detection of the state of each atom relies on the lossof atoms in the Rydberg state, which are not recaptured inthe optical tweezers. This detection method is thus subject tosmall detection errors [46]. In particular, an atom actually in|↓〉 has a small probability ε to be lost and thus incorrectlyinferred to be in |↑〉. As in [30], we measure ε in a calibrationexperiment, and then include the effect of detection errors onall the theoretical curves.

An antiferromagnetic phase is expected when interactionbetween parallel spins are repulsive, corresponding to the casewhere U > 0 in the Hamiltonian Eq. (1). Fig. 6 (a) shows the

Page 9: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

9

Figure U/h Ωmax/(2π) δfinal/(2π) trise tsweep tfall

(MHz) (MHz) (MHz) (µs) (µs) (µs)

2 1.0 2.3 [−2, 6] 0.25δfinal − δ0

2π · 10 (MHz)0.5

3 2.7 1.8 4.5 0.25 [0.1, 1.3] 0.25

4 2.7 1.8 4.5 0.25 0.44 0.25

5(a) 2.7 1.8 4.5 0.25 0.44 0.25

5(b) 0.8 0.6 1.6 0.25 5.5 0.25

TABLE I. Experimental parameters used for the data presented in themain text.

energy levels for a 2×2 square matrix of atoms as a functionof the detuning δ, with U > 0 and ~Ω ' 0.5U . The lowestenergy configuration with two |↑〉 spins is then an antiferro-magnetic state for 0 < ~δ < 2U . Then, starting from a detun-ing δ < 0 with all atoms in |↓〉, and ramping it up to δ ≈ U/~(dashed arrow), the lowest energy state is evolving adiabati-cally to the antiferromagnetic state. In our experiment, the in-teraction between |↑〉 =

∣∣64D3/2,mj = 3/2⟩

spin states arein fact attractive (U < 0). Consequently, the antiferromag-netic state is never the lowest energy configuration whateverthe value of δ, see Fig. 6 (b). Nevertheless, starting from adetuning δ > 0 and ramping it down to δ ≈ U/~, the sys-tem evolves from the initial state to an antiferromagnetic statewhile following the highest-energy level. In this case, prepar-ing the antiferromagnetic state actually means obtaining theground state of −H , hence the change of sign of the detun-ings needed to reach the correlated phase. For our parameters,the van der Waals interaction is not only attractive, but it isalso slightly anisotropic [25, 34, 52], being about three timesas small in the horizontal direction as in the vertical one (alongz). We compensated for this difference by slightly distortingthe arrays along z by a factor ∼ 31/6.

Appendix B: Description of the “order-by-disorder” process

As mentioned in Sec. III, the triangular lattice presents aninteresting feature for ~δ/U = z/2: there the model is fullyfrustrated in the classical limit; the interactions on the bondsof each single triangle of the lattice cannot be simultane-ously satisfied under the given density constraint n = 1/2leading to an extensive ground-state degeneracy [53]. Uponswitching on the transverse field, this degeneracy is liftedby a process called “order-by-disorder” (OBD), i.e. the ad-dition of “disorder” (here the quantum fluctuations providedby the transverse field Ω) selects a subset of configurationsdisplaying an ordered pattern. In the case under consider-ation here, OBD leads to the selection of a 3-sublattice or-

0 1 2 3 4

~δ/U

0.0

0.2

0.4

0.6

0.8

1.0

n

1

22

2

1

2

Square N = 36

OBC

PBC

0 1 2 3 4 5 6

~δ/U

1

2812

450

531

1

3

3

Triangular N = 36

FIG. 7. Classical (Ω = 0) phase diagrams for arrays with openboundary conditions. Rydberg density n as a function of x = ~δ/Ufor the N = 36 square and triangular arrays used in the experiment.Blue lines correspond to open boundary conditions while red dashedlines correspond to periodic boundary conditions (bulk phase dia-gram). The numbers indicate the degeneracy of the classical config-urations for the distinct plateaus.

dered phase with Rydberg occupations of the three sublattices(nA, nB , nC) = (λ, 1 − λ, 1/2), λ 6= 1/2. This phase laterundergoes a quantum phase transition in the 3d-XY universal-ity class to the paramagnetic phase for a larger Ω [54, 55].

Appendix C: Finite-size effects

The two-dimensional arrays used here present open bound-ary conditions (OBC). The lattice sites along the boundaryhave less neighbors than the sites in the bulk and therefore ex-perience less interactions. This modifies the “phase diagram”with respect to the bulk phase diagrams presented in Fig. 1. InFig. 7 we show the classical (i.e. Ω = 0) ground-state config-urations for the N = 36 square and triangular arrays used inthe experiment and compare them to the bulk phase diagrams.The sites at the boundary are more easily excited to the Ry-dberg state when δ increases, leading to a larger number ofdensity plateaus. We have checked that for the parametersused in this paper the OBC and bulk phases feature the sametypical short-range correlations.

The Neel structure factor SNeel shows almost no finite-sizeeffects for short-range ordered systems as long as the systemsizes are larger than the correlation length. For finite arraysthe number of pairs of sites with larger distances |k| + |l| isstrongly reduced leading to larger errors for the connected cor-relations g(2)(k, l) due to reduced statistics. In order to keepreasonable values and errors for the Neel structure factor wehave restricted the summation to |k| + |l| ≤ 4 in the experi-mental evaluation of SNeel.

Appendix D: Numerical Methods

We perform numerical exact diagonalization simulations ofthe time evolution on medium-size lattices for both Hamilto-nian systems without any dissipation and the same systemswith additional dephasing terms in the framework of a mas-ter equation in Lindblad form. In both cases we construct the

Page 10: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

10

complete many-body Hilbert space of the system on the fi-nite lattice and do not perform any truncation to calculate thetime evolution. Therefore, during the time evolution the sys-tem can explore the entire Hilbert space and all phases in thephase diagram can potentially be obtained.

1. Unitary time evolution

To simulate the unitary time evolution of a time-dependentHamiltonian H(t), we approximate H(t) by a piecewise con-stant Hamiltonian Happrox(t) on nsteps intervals of length ∆tsuch that

Happrox(t) =H(n∆t) +H((n+ 1)∆t)

2, (D1)

for n∆t ≤ t < (n + 1)∆t. We then compute the evolutionoperator within each interval with a Krylov-type matrix expo-nential approach [56] and use the evolved state of the previousinterval as starting state. We choose nsteps such that the resultsdo not differ from a simulation with nsteps/2 intervals withina demanded accuracy. In most of the presented simulationsnsteps = 200.

We are able to calculate the unitary time evolution of sys-tem sizes of up to about 30 lattice sites with a reasonableamount of resources. In the present study most of the numeri-cal results have been obtained on 4×4 and 5×5 square lattices.Due to the relatively short correlation lengths observed in theexperiments, the numerical results on the smaller systems cannevertheless be used to model the observed correlation func-tions on the larger, experimentally realizable lattices.

2. Time evolution in the presence of dephasing

The experimental system features several sources of imper-fections [46] and the description of its evolution as a pureHamiltonian is not exact. In particular, phase noise of theexcitation lasers and Doppler shifts lead to dephasing, alreadyfor a single particle. We thus perform simulations with a phe-nomenological local dephasing model with a rate γ obtainedby fitting experimental single atom Rabi oscillations. We ob-tain dephasing rates ~γ/U ≈ 1.1 − 1.4 for the data shown inthis paper. The time evolution of the density matrix ρ(t) of themany-body system is then described by the following masterequation in Lindblad form

d

dtρ = − i

~[H, ρ] + L[ρ] , (D2)

with a Liouvillian

L[ρ] =∑i

γ

2(2niρni − niρ− ρni) . (D3)

Direct simulations of the master equation are only possiblefor small lattices as the memory demand for saving densitymatrices grows as O(4N ) for the considered Hilbert space onN sites. Thus we use a Monte Carlo wave-function (MCWF)

FIG. 8. Illustration of the Lieb-Robinson bounds. “Significant” cor-relations spread along a light-cone d = 2 v t. Along and inside thelight-cone (grey shaded area) correlations of orderO(1) can develop.Outside the light-cone d > 2 v t correlations are immediately builtup for each t > 0 but are exponentially suppressed in d− 2 v t.

method [57–59] where only a single wave-function has to bestored such that one can simulate system sizes of up to aroundtwenty sites. The MCWF method evolves a starting state withan effective non-hermitian Hamiltonian; at each time step aquantum jump corresponding to the given dissipation operatorcollapses the evolved state with a given probability. Averagingover many such quantum trajectories —we use 1280 for thepresented results— allows reconstructing the density matrixand thus computing any observable during the time evolution.In practice, to perform simulations with dissipation, we usethe Python toolbox “QuTiP” [60].

Appendix E: Lieb-Robinson bounds

In non-relativistic quantum mechanics, one might believethat information propagation is instantaneous because there isno explicit speed of light limiting the propagation. However,in 1972 Lieb and Robinson [3] proved that in an extendedquantum many-body system with a finite-dimensional localHilbert space and sufficiently local interactions, there is nev-ertheless a characteristic velocity emerging, which defines anapproximate light-cone for the propagation of information andimplementing causality.

In Ref. [4] the Lieb-Robinson bound has been generalizedto equal-time connected correlation functions of two opera-tors (normalized to unity) acting on spatial regions A and B(of size |A| and |B|) at a distance d apart from each other aftersome time evolution of duration t and starting from an initialstate with exponentially decaying correlations with a corre-lation length χ. Then the connected equal-time correlationg(2)(d, t) := 〈OA(t)OB(t)〉−〈OA(t)〉〈OB(t)〉 is bounded asfollows:

|g(2)(d, t)| ≤ c(|A|+ |B|) exp[−(d− 2vt)/χ′], (E1)

with χ′ = χ + 2ζ. The coefficients c, v and ζ depend onthe Hamiltonian and the considered operators OA and OB .

Page 11: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

11

FIG. 9. Illustration of the different paths connecting two sites with adistance (k, l) contribution to the leading-order short-time expansionon the (a) square lattice and (b) triangular lattice. The number inbrackets shows the binomial coefficient C|k|+|l||k−l| (assuming k, l ≥ 0)corresponding to the number of distinct paths. The arrows show thepotentially different couplings along paths contributing to the short-time expansion. The grey lines exemplary show the three distinctpaths contributing to the correlation indicated by a grey square.

The velocity v is particularly important and is called the Lieb-Robinson velocity.

The importance of this result for the present study is thatcorrelations at a distance d are exponentially suppressed in(d− 2vt)/χ′ as long as the time t < d/(2v). After that time,the correlations are bounded by a number of O(1). This timedependence is visualized in Fig. 8. In closing we note thatthe theorem proves rigorous theoretical bounds, however theactual dynamics does not necessarily saturate these bounds.

Appendix F: Short-time expansion

In this appendix we provide more details on the short-timeexpansion of the connected correlation function g(2).

1. Theoretical description

In order to simplify the notation we use the nearest-neighbor Hamiltonian (~ = 1):

H(t) =∑〈i,j〉

Uij ninj − δ(t)∑i

ni + Ω/2∑i

σxi . (F1)

Here, we keep the possibility that nearest-neighbor interac-tions could be different from atom to atom. We considersimple ramp shapes characterized by a constant Ω, a dura-tion T and a linear δ(t)-dependence between δ(0) = δ0 andδ(T ) = δfinal.

We are interested in the limit of very short duration Tand in particular how the correlations at various distancesbuild up as we vary the duration T . Formally, we com-pute the full many-body propagator U(T ) which solves thetime-dependent Schrodinger equation for the time-dependentHamiltonian Eq. (F1), allowing to determine the many-bodywave function at any time T via |ψ(T )〉 = U(T )|ψ(0)〉. Wethen express the connected correlation function as

g(2)[(k, l)](T ) = 〈ψ(T )|n(0,0)n(k,l)|ψ(T )〉 (F2)− 〈ψ(T )|n(0,0)|ψ(T )〉〈ψ(T )|n(k,l)|ψ(T )〉

For the linear ramps considered here, a Magnus expan-sion [61] of the propagator is appropriate. We rely on theleading Magnus expansion term, which can be written as

U(T ) ≈ exp[−iTHavg] , (F3)

with

Havg =1

T

∫ T

0

H(t′)dt′ (F4)

For the ramp shape considered here, Havg is independent of Tand is of the same form as (F1) with Uij and Ω unchanged,while δ(t) is replaced by δavg = (δ0 + δfinal)/2.

We now calculate symbolic expressions for the power seriesin T of the connected correlators (F2) relying on the propaga-tor in the leading Magnus expansion form and starting froman initial state with all sites in the atomic ground state. When(UT/~,ΩT, δT 1), we find the following expressions forthe leading order in T for a single path on a lattice linkingsites (0, 0) and (k, l) (for the sake of simplicity the results areonly given for Uij = U ):

• nearest neighbor correlation |k|+ |l| = 1

g(2)(T ) = − 1

288

(U2 − 3Uδavg

)Ω4T 6, (F5)

• next nearest neighbor correlation |k|+ |l| = 2

g(2)(T ) =Ω6T 10

2419200

(77U4 − 340U3δavg + 375U2δ2

avg

),

(F6)

• third nearest neighbor correlation |k|+ |l| = 3

g(2)(T ) = − Ω8T 14

26824089600

(4279U6 − 24766U5δavg

+46725U4δ2avg − 28350U3δ3

avg

). (F7)

While the expressions look complicated, a common struc-ture emerges: the leading T dependence for a Manhattan dis-tance m = |k|+ |l| is proportional to T 2+4m. Neglecting the

Page 12: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

12

−4

−2

0

2

4

l

2

3

4

1

3

4

1

2

4

3

3

4

4

4

2

3

4

2

−2.5 0.0 2.5k

−4

−2

0

2

4

l

−0.08

−0.06

−0.04

−0.02

0.00

0.02

0.04

0.06

0.08

4×g

(2) (k,l

)

−0.6

−0.4

−0.2

0.0

0.2

0.4

0.6

4×g

(2) (k,l

)

(a)

(b)

FIG. 10. Correlations on the N = 36-site triangular array (only thefirst four Manhattan shells are shown). (a) Experimentally measuredcorrelations after a sweep on the triangular array featuring a pro-nounced Manhattan shell structure g(2)(k, l) ∝ (−1)m. The num-bers give the value of m = |k| + |l|. (b) Correlations of the equalsuperposition of the classical groundstates for δfinal as reached at theend of the sweep used in (a). These correlations are in qualitativeagreement with the correlations expected in the 1/3 AF state in theright panel of Fig. 1(c).

actual value of the coefficients for the moment in an admit-tedly crude approximation, these results suggest that correla-tions at a next shell are suppressed by a factor T 4 comparedto the previous shell, suggesting an exponential spatial decayof the connected correlations at short times with m. Theseresults also provide an insight as to why more distant shellsrequire longer times to develop appreciable correlations: theyrequire larger values of T to overcome the initial high powersin T suppression of the correlations.

Importantly, the neglected higher-order terms in the Mag-nus expansions do not alter the leading order in T for the

considered ramp shapes, but only affect the sub-leading T -coefficients. Finally, we note that at very short times, cor-relations induced by the direct van der Waals tail of the in-teractions compete with the high-order nearest-neighbor con-siderations here. Although strictly speaking this leads to abreakdown of the exponential suppression of correlations be-yond the light-cone, the smallness of this effect prevents itsobservation in the present experiments.

2. Lattice embedding

The expressions derived above are lattice independent, aslong as there is a chain of m (m = |k| + |l| above) succes-sive nearest neighbor interactions linking the two consideredsites [62]. These expressions can now be embedded in anylattice (e.g. cubic, kagome, honeycomb, ...), and the actualcoefficients of the short-time series can be derived by deter-mining the corresponding embedding coefficients. The em-bedding counting is illustrated in Fig. 9 for the square andthe triangular lattice, yielding binomial coefficients multiply-ing the symbolic expressions for a single path derived above.Note that for the case Uz 6= Uw the coefficients of the singlepaths can already differ before taking the embedding factorsinto account.

Appendix G: Correlations on the triangular lattice: Short-timeversus ground state correlations

The short-time expansion on the square lattice produces thesame staggered correlation pattern as the one of the ground-state of an AF ordered phase. There is thus a similarity be-tween the sign structure of the correlations at short times andat very long times (i.e. the adiabatic limit).

On the triangular lattice, however, the correlations obtainedin the experiment [see Fig. 10(a)], which are in good agree-ment with the short-time expansion as shown in Fig. 5(c-d),differ significantly from the ones expected for the AF groundstate with density 1/3 [Fig. 10(b)]. The origin of this discrep-ancy may be related to the frustrated character of the triangulargeometry and will be the subject of future work.

[1] R. Nandkishore and D. A. Huse, “Many-body localization andthermalization in quantum statistical mechanics,” Annu. Rev.Condens. Matter Phys. 6, 15 (2015).

[2] E. Altman and R. Vosk, “Universal Dynamics and Renormaliza-tion in Many-Body-Localized Systems,” Annu. Rev. Condens.Matter Phys. 6, 383 (2015).

[3] E. H. Lieb and D. W. Robinson, “The finite group velocity ofquantum spin systems,” Comm. Math. Phys. 28, 251 (1972).

[4] S. Bravyi, M. B. Hastings, and F. Verstraete, “Lieb-RobinsonBounds and the Generation of Correlations and TopologicalQuantum Order,” Phys. Rev. Lett. 97, 050401 (2006).

[5] J. Eisert, M. van den Worm, S. R. Manmana, and M. Kast-ner, “Breakdown of quasilocality in long-range quantum lattice

models,” Phys. Rev. Lett. 111, 260401 (2013).[6] M. Foss-Feig, Z.-X. Gong, C. W. Clark, and A. V. Gorshkov,

“Nearly linear light cones in long-range interacting quantumsystems,” Phys. Rev. Lett. 114, 157201 (2015).

[7] I. G. White, B. Sundar, and K. R. A. Hazzard, “Quan-tum dynamics from a numerical linked cluster expansion,”arXiv:1710.07696 (2017).

[8] L. Cevolani, G. Carleo, and L. Sanchez-Palencia, “Spreadingof correlations in exactly solvable quantum models with long-range interactions in arbitrary dimensions,” New J. Phys. 18,093002 (2016).

[9] I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simula-tion,” Rev. Mod. Phys. 86, 153 (2014).

Page 13: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

13

[10] K. Kim, M.-S. Chang, S. Korenblit, R. Islam, E. E. Edwards,J. K. Freericks, G.-D. Lin, L.-M. Duan, and C. Monroe, “Quan-tum simulation of frustrated Ising spins with trapped ions,” Na-ture 465, 590 (2010).

[11] P. Richerme, C. Senko, J. Smith, A. Lee, S. Korenblit, andC. Monroe, “Experimental performance of a quantum simula-tor: Optimizing adiabatic evolution and identifying many-bodyground states,” Phys. Rev. A 88, 012334 (2013).

[12] J. Simon, W. Bakr, R. Ma, M. Tai, P. Preiss, and M. Greiner,“Quantum simulation of antiferromagnetic spin chains in an op-tical lattice,” Nature 472, 307 (2011).

[13] J. Struck, C. Olschlager, R. Le Targat, P. Soltan-Panahi,A. Eckardt, M. Lewenstein, P. Windpassinger, and K. Seng-stock, “Quantum simulation of frustrated classical magnetismin triangular optical lattices,” Science 333, 996 (2011).

[14] R. Landig, L. Hruby, N. Dogra, M. Landini, R. Mottl, T. Don-ner, and T. Esslinger, “Quantum phases from competing short-and long-range interactions in an optical lattice,” Nature 532,476 (2016).

[15] L. W. Cheuk, M. A. Nichols, K. R. Lawrence, M. Okan,H. Zhang, E. Khatami, N. Trivedi, T. Paiva, M. Rigol, andM. W. Zwierlein, “Observation of spatial charge and spin cor-relations in the 2d Fermi-Hubbard model,” Science 353, 1260(2016).

[16] P. T. Brown, D. Mitra, E. Guardado-Sanchez, P. Schauß, S. S.Kondov, E. Khatami, T. Paiva, N. Trivedi, D. A. Huse, andW. S. Bakr, “Spin-imbalance in a 2d Fermi-Hubbard system,”Science 357, 1385 (2017).

[17] A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M. Kansz-Nagy, R. Schmidt, F. Grusdt, E. Demler, D. Greif, andM. Greiner, “A cold-atom Fermi-Hubbard antiferromagnet,”Nature 545, 462 (2017).

[18] M. Cheneau, P. Barmettler, D. Poletti, M. Endres, P. Schausz,T. Fukuhara, C. Gross, I. Bloch, C. Kollath, and S. Kuhr,“Light-cone-like spreading of correlations in a quantum many-body system,” Nature 481, 484 (2012).

[19] A. de Paz, A. Sharma, A. Chotia, E. Marechal, J. H. Huckans,P. Pedri, L. Santos, O. Gorceix, L. Vernac, and B. Laburthe-Tolra, “Nonequilibrium quantum magnetism in a dipolar latticegas,” Phys. Rev. Lett. 111, 185305 (2013).

[20] P. Richerme, Z.-X. Gong, A. Lee, C. Senko, J. Smith, M. Foss-Feig, S. Michalakis, A. V. Gorshkov, and C. Monroe, “Non-local propagation of correlations in quantum systems with long-range interactions,” Nature 511, 198 (2014).

[21] S. Hild, T. Fukuhara, P. Schauß, J. Zeiher, M. Knap, E. Demler,I. Bloch, and C. Gross, “Far-from-equilibrium spin transportin Heisenberg quantum magnets,” Phys. Rev. Lett. 113, 147205(2014).

[22] F. Meinert, M. J. Mark, E. Kirilov, K. Lauber, P. Weinmann,M. Grobner, A. J. Daley, and H.-C. Nagerl, “Observation ofmany-body dynamics in long-range tunneling after a quantumquench,” Science 344, 1259 (2014).

[23] P. Jurcevic, B. P. Lanyon, P. Hauke, C. Hempel, P. Zoller,R. Blatt, and C. F. Roos, “Quasiparticle engineering and entan-glement propagation in a quantum many-body system,” Nature511, 202 (2014).

[24] J. Zhang, G. Pagano, P. W. Hess, A. Kyprianidis, P. Becker,H. Kaplan, A. V. Gorshkov, Z. X. Gong, and C. Monroe, “Ob-servation of a Many-Body Dynamical Phase Transition with a53-Qubit Quantum Simulator,” arXiv:1708.01044 (2017).

[25] A. Browaeys, D. Barredo, and T. Lahaye, “Experimental inves-tigations of dipole-dipole interactions between a few Rydbergatoms,” J. Phys. B 49, 152001 (2016).

[26] P. Schauß, M. Cheneau, M. Endres, T. Fukuhara, S. Hild,

A. Omran, T. Pohl, C. Gross, S. Kuhr, and I. Bloch, “Observa-tion of mesoscopic crystalline structures in a two-dimensionalRydberg gas,” Nature 491, 87 (2012).

[27] P. Schauss, J. Zeiher, T. Fukuhara, S. Hild, M. Cheneau,T. Macri, T. Pohl, I. Bloch, and C. Gross, “Crystallization inIsing quantum magnets,” Science 347, 1455 (2015).

[28] T. Pohl, E. Demler, and M. D. Lukin, “Dynamical crystalliza-tion in the dipole blockade of ultracold atoms,” Phys. Rev. Lett.104, 043002 (2010).

[29] D. Petrosyan, K. Mølmer, and M. Fleischhauer, “On the adi-abatic preparation of spatially-ordered Rydberg excitations ofatoms in a one-dimensional optical lattice by laser frequencysweeps,” J. Phys. B 49, 084003 (2016).

[30] H. Labuhn, D. Barredo, S. Ravets, S. de Leseleuc, T. Macrı,T. Lahaye, and A. Browaeys, “Tunable two-dimensional arraysof single Rydberg atoms for realizing quantum Ising models,”Nature 534, 667 (2016).

[31] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran,H. Pichler, S. Choi, A. S. Zibrov, M. Endres, M. Greiner,V. Vuletic, and M. D. Lukin, “Probing many-body dynamicson a 51-atom quantum simulator,” arXiv:1707.04344 (2017).

[32] D. Barredo, S. de Leseleuc, V. Lienhard, T. Lahaye, andA. Browaeys, “An atom-by-atom assembler of defect-free ar-bitrary 2d atomic arrays,” Science 354, 1021 (2016).

[33] The next-nearest neighbor interaction strength is (√

2)6 = 8times smaller than the NN one and is thus one order of mag-nitude smaller than the Rabi frequency drive Ω and the inverseexperimental time scale 1/ttot.

[34] S. de Leseleuc, S. Weber, V. Lienhard, D. Barredo, H. P.Buchler, T. Lahaye, and A. Browaeys, “Accurate mapping ofmultilevel Rydberg atoms on interacting spin-1/2 particles forthe quantum simulation of Ising models,” arXiv:1710.06156(2017).

[35] S. Sachdev, Quantum Phase Transitions, 2nd ed. (CambridgeUniversity Press, 2011).

[36] H. W. J. Blote and Y. Deng, “Cluster Monte Carlo simulation ofthe transverse Ising model,” Phys. Rev. E 66, 066110 (2002).

[37] A. Das and B. K. Chakrabarti, “Colloquium: Quantum anneal-ing and analog quantum computation,” Rev. Mod. Phys. 80,1061 (2008).

[38] M. Schuler, S. Whitsitt, L.-P. Henry, S. Sachdev, and A. M.Lauchli, “Universal signatures of quantum critical points fromfinite-size torus spectra: A window into the operator content ofhigher-dimensional conformal field theories,” Phys. Rev. Lett.117, 210401 (2016).

[39] C. Laumann, R. Moessner, A. Scardicchio, and S. Sondhi,“Quantum annealing: The fastest route to quantum computa-tion?” Eur. Phys. J. ST 224, 75 (2015).

[40] In order to improve the statistics we symmetrize the experimen-tal data for a given (k, l) over the four quadrants (±k,±l) forthe square lattice (and in an adapted way for the triangular lat-tice). Since these operations are symmetries of the setup, thesymmetrization does not alter genuine features in the experi-mental data.

[41] Using the 1/r6 interaction would lead to finer structures aroundx = 0 and x = 4, but with very narrow steps in x.

[42] In the bulk or for L even, the AF state is the equal superposi-tion of the two classical Neel states. It has a uniform Rydbergdensity ni = 1/2, while g(2)(k, l) = (−1)|k|+|l| × 1/4. Instark contrast, on open boundary L × L arrays with L odd, asingle Neel ground state is selected. This state exhibits a stag-gered pattern of Rydberg densities, while the g(2) correlationsvanish identically.

[43] We have checked that using NN interactions instead of the exact

Page 14: Laboratoire Charles Fabry, Institut d’Optique Graduate ... · PDF fileLaboratoire Charles Fabry, Institut d’Optique Graduate School, ... France Michael Schuler , Louis-Paul Henry,

14

van der Waals coupling does not lead to appreciable changes inthe dynamics.

[44] A. M. Lauchli and C. Kollath, “Spreading of correlations andentanglement after a quench in the one-dimensional Bose-Hubbard model,” J. Stat. Mech. Theor. Exp. 2008, P05018(2008).

[45] T. Langen, R. Geiger, M. Kuhnert, B. Rauer, and J. Schmied-mayer, “Local emergence of thermal correlations in an isolatedquantum many-body system,” Nat. Phys. 9, 640 (2013).

[46] S. de Leseleuc et al., in preparation.[47] S. Humeniuk, “Quantum Monte Carlo study of long-range

transverse-field Ising models on the triangular lattice,” Phys.Rev. B 93, 104412 (2016).

[48] A. M. Lauchli and R. Moessner, “Quantum simulations madeeasy plane,” arXiv:1504.04380 (2015).

[49] D. Barredo, H. Labuhn, S. Ravets, T. Lahaye, A. Browaeys,and C. S. Adams, “Coherent excitation transfer in a spin chainof three Rydberg atoms,” Phys. Rev. Lett. 114, 113002 (2015).

[50] S. de Leseleuc, D. Barredo, V. Lienhard, A. Browaeys, andT. Lahaye, “Optical control of the resonant dipole-dipole inter-action between Rydberg atoms,” Phys. Rev. Lett. 119, 053202(2017).

[51] D. Peter, S. Muller, S. Wessel, and H. P. Buchler, “Anomalousbehavior of spin systems with dipolar interactions,” Phys. Rev.Lett. 109, 025303 (2012).

[52] D. Barredo, S. Ravets, H. Labuhn, L. Beguin, A. Vernier, F. No-grette, T. Lahaye, and A. Browaeys, “Demonstration of a strong

Rydberg blockade in three-atom systems with anisotropic inter-actions,” Phys. Rev. Lett. 112, 183002 (2014).

[53] G. H. Wannier, “Antiferromagnetism. The triangular Ising net,”Phys. Rev. 79, 357 (1950).

[54] R. Moessner, S. L. Sondhi, and P. Chandra, “Two-DimensionalPeriodic Frustrated Ising Models in a Transverse Field,” Phys.Rev. Lett. 84, 4457 (2000).

[55] R. Moessner and S. L. Sondhi, “Ising models of quantum frus-tration,” Phys. Rev. B 63, 224401 (2001).

[56] T. J. Park and J. C. Light, “Unitary quantum time evolution byiterative Lanczos reduction,” J. Chem. Phys. 85, 5870 (1986).

[57] J. Dalibard, Y. Castin, and K. Mølmer, “Wave-function ap-proach to dissipative processes in quantum optics,” Phys. Rev.Lett. 68, 580 (1992).

[58] R. Dum, P. Zoller, and H. Ritsch, “Monte Carlo simulation ofthe atomic master equation for spontaneous emission,” Phys.Rev. A 45, 4879 (1992).

[59] K. Mølmer, Y. Castin, and J. Dalibard, “Monte Carlo wave-function method in quantum optics,” J. Opt. Soc. Am. B 10,524 (1993).

[60] J. Johansson, P. Nation, and F. Nori, “QuTiP 2: A Pythonframework for the dynamics of open quantum systems,” Com-put. Phys. Commun. 184, 1234 (2013).

[61] S. Blanes, F. Casas, J. Oteo, and J. Ros, “The Magnus expan-sion and some of its applications,” Phys. Rep. 470, 151 (2009).

[62] J. Oitmaa, C. Hamer, and W. Zheng, Series Expansion Methodsfor Strongly Interacting Lattice Models (Cambridge UniversityPress, 2006).