Short-Range Quantum Magnetism of Ultracold Fermions in an Optical Lattice

10
Short-range quantum magnetism of ultracold fermions in an optical lattice Daniel Greif, 1 Thomas Uehlinger, 1 Gregor Jotzu, 1 Leticia Tarruell, 1,2, * and Tilman Esslinger 1, 1 Institute for Quantum Electronics, ETH Zurich, 8093 Zurich, Switzerland 2 LP2N UMR 5298, Univ. Bordeaux 1, Institut d’Optique and CNRS, 351 cours de la Lib´ eration, 33405 Talence, France The exchange coupling between quantum mechanical spins lies at the origin of quantum magnetism. We report on the observation of nearest-neighbor magnetic spin correlations emerging in the many-body state of a thermalized Fermi gas in an optical lattice. The key to obtaining short-range magnetic order is a local redistri- bution of entropy within the lattice structure. This is achieved in a tunable-geometry optical lattice, which also enables the detection of the magnetic correlations. We load a low-temperature two-component Fermi gas with repulsive interactions into either a dimerized or an anisotropic simple cubic lattice. For both systems the corre- lations manifest as an excess number of singlets as compared to triplets consisting of two atoms with opposite spins. For the anisotropic lattice, we determine the transverse spin correlator from the singlet-triplet imbalance and observe antiferromagnetic correlations along one spatial axis. Our work paves the way for addressing open problems in quantum magnetism using ultracold fermions in optical lattices as quantum simulators. Quantum magnetism describes quantum many-body states of spins coupled by exchange interactions and lies at the heart of many fundamental phenomena in condensed mat- ter physics 1,2 . While spin systems often tend to show long-range order at low temperatures, the fascinating in- terplay of exchange interactions with geometry and quan- tum fluctuations can lead to quantum states characterized by their short-range magnetic order. Examples include valence- bond crystals, spin-liquids and possibly high-temperature superconductors 1–4 . Quite remarkably, the underlying many- body physics gives rise to computationally and theoretically intractable regimes even in the phase diagrams of simple mod- els, such as the Fermi-Hubbard model. Moreover, the direct measurement of local spin correlations in solids remains a ma- jor challenge. The controlled setting of ultracold fermionic atoms in op- tical lattices is regarded as a promising route to gain new in- sights into phases driven by quantum magnetism 5–7 . This ap- proach offers experimental access to a clean and highly flex- ible Fermi-Hubbard model with a unique set of observables 8 . For repulsively interacting fermions density ordering in the metal-Mott insulator transition could already be explored experimentally 9,10 . Yet, progress towards entering the regime of quantum magnetism has been hindered by the ultra-low temperatures, and entropies, required to observe exchange- driven spin ordering in optical lattices. For bosonic quantum gases promising developments have been reported: by map- ping the spin to other degrees of freedom the temperature lim- itation could be circumvented, which allowed the exploration of one-dimensional decoupled Ising spin chains and the simu- lation of classical magnetism on a triangular lattice 11,12 . Fur- thermore, exchange interactions were observed in artificially prepared arrays of isolated double-wells or plaquettes 13,14 . To overcome these limitations and directly access exchange-driven physics in thermalized systems, cooling schemes based on the redistribution of entropy between differ- ent regions of the trap have been suggested 15,16 . In this work, we instead propose and implement a local entropy redistribu- tion scheme within the lattice structure to reach the regime of quantum magnetism. The atoms are either prepared in a dimerized or an anisotropic simple cubic lattice, see Fig. 1. In both geometries, a subset of links of the underlying simple cu- bic lattice is set to a larger exchange energy as compared to the other links. As a result, the entropy is predominantly stored in configurations involving the weak links. For fixed total en- tropy in the trapped system, this essentially allows us to reach temperatures between the two exchange energy scales. In the dimerized lattice the resulting correlations on the strong links correspond to an excess population of the low energy singlet as compared to the triplet state – in close resemblance to an explicit valence-bond crystal in the Heisenberg model 17 . In the anisotropic lattice the low temperatures lead to antiferro- magnetic spin correlations along one spatial axis, the trans- verse component of which is also detected via a singlet-triplet imbalance. For both systems we study the dependence of the spin correlations as a function of temperature and tunneling balance and find good agreement with theory. a b Dimerized cubic lattice Anisotropic cubic lattice t t t t t t d t s t s x y z Singlet Triplet Antiferromagnetic spin correlation FIG. 1. Magnetic spin correlations. Schematic view of the nearest-neighbor spin correlations observed in the experiment. A two-component mixture of fermionic atoms (red and blue) is pre- pared close to half-filling in a cubic lattice with two different tunnel coupling configurations. a, Dimerized lattice with the strong dimer links t d and weaker links t. Low temperatures lead to an excess number of singlets over triplets. b, Anisotropic lattice with strong and weak tunneling ts and t along different spatial axes. Antifer- romagnetic spin correlations in the transverse direction are formed along the strong link direction. In both figures exemplary thermal excitations in the form of spin excitations or holes are shown. arXiv:1212.2634v2 [cond-mat.quant-gas] 27 Jun 2013

Transcript of Short-Range Quantum Magnetism of Ultracold Fermions in an Optical Lattice

Short-range quantum magnetismof ultracold fermions in an optical lattice

Daniel Greif,1 Thomas Uehlinger,1 Gregor Jotzu,1 Leticia Tarruell,1, 2, ∗ and Tilman Esslinger1, †

1Institute for Quantum Electronics, ETH Zurich, 8093 Zurich, Switzerland2LP2N UMR 5298, Univ. Bordeaux 1, Institut d’Optique and CNRS, 351 cours de la Liberation, 33405 Talence, France

The exchange coupling between quantum mechanical spins lies at the origin of quantum magnetism. Wereport on the observation of nearest-neighbor magnetic spin correlations emerging in the many-body state of athermalized Fermi gas in an optical lattice. The key to obtaining short-range magnetic order is a local redistri-bution of entropy within the lattice structure. This is achieved in a tunable-geometry optical lattice, which alsoenables the detection of the magnetic correlations. We load a low-temperature two-component Fermi gas withrepulsive interactions into either a dimerized or an anisotropic simple cubic lattice. For both systems the corre-lations manifest as an excess number of singlets as compared to triplets consisting of two atoms with oppositespins. For the anisotropic lattice, we determine the transverse spin correlator from the singlet-triplet imbalanceand observe antiferromagnetic correlations along one spatial axis. Our work paves the way for addressing openproblems in quantum magnetism using ultracold fermions in optical lattices as quantum simulators.

Quantum magnetism describes quantum many-body statesof spins coupled by exchange interactions and lies at theheart of many fundamental phenomena in condensed mat-ter physics1,2. While spin systems often tend to showlong-range order at low temperatures, the fascinating in-terplay of exchange interactions with geometry and quan-tum fluctuations can lead to quantum states characterized bytheir short-range magnetic order. Examples include valence-bond crystals, spin-liquids and possibly high-temperaturesuperconductors1–4. Quite remarkably, the underlying many-body physics gives rise to computationally and theoreticallyintractable regimes even in the phase diagrams of simple mod-els, such as the Fermi-Hubbard model. Moreover, the directmeasurement of local spin correlations in solids remains a ma-jor challenge.

The controlled setting of ultracold fermionic atoms in op-tical lattices is regarded as a promising route to gain new in-sights into phases driven by quantum magnetism5–7. This ap-proach offers experimental access to a clean and highly flex-ible Fermi-Hubbard model with a unique set of observables8.For repulsively interacting fermions density ordering in themetal-Mott insulator transition could already be exploredexperimentally9,10. Yet, progress towards entering the regimeof quantum magnetism has been hindered by the ultra-lowtemperatures, and entropies, required to observe exchange-driven spin ordering in optical lattices. For bosonic quantumgases promising developments have been reported: by map-ping the spin to other degrees of freedom the temperature lim-itation could be circumvented, which allowed the explorationof one-dimensional decoupled Ising spin chains and the simu-lation of classical magnetism on a triangular lattice11,12. Fur-thermore, exchange interactions were observed in artificiallyprepared arrays of isolated double-wells or plaquettes13,14.

To overcome these limitations and directly accessexchange-driven physics in thermalized systems, coolingschemes based on the redistribution of entropy between differ-ent regions of the trap have been suggested15,16. In this work,we instead propose and implement a local entropy redistribu-tion scheme within the lattice structure to reach the regimeof quantum magnetism. The atoms are either prepared in a

dimerized or an anisotropic simple cubic lattice, see Fig. 1. Inboth geometries, a subset of links of the underlying simple cu-bic lattice is set to a larger exchange energy as compared to theother links. As a result, the entropy is predominantly storedin configurations involving the weak links. For fixed total en-tropy in the trapped system, this essentially allows us to reachtemperatures between the two exchange energy scales. In thedimerized lattice the resulting correlations on the strong linkscorrespond to an excess population of the low energy singletas compared to the triplet state – in close resemblance to anexplicit valence-bond crystal in the Heisenberg model17. Inthe anisotropic lattice the low temperatures lead to antiferro-magnetic spin correlations along one spatial axis, the trans-verse component of which is also detected via a singlet-tripletimbalance. For both systems we study the dependence of thespin correlations as a function of temperature and tunnelingbalance and find good agreement with theory.

a b

Dimerized cubic lattice Anisotropic cubic lattice

t

t

t

t t

td ts ts

xy

z

Singlet Triplet Antiferromagneticspin correlation

FIG. 1. Magnetic spin correlations. Schematic view of thenearest-neighbor spin correlations observed in the experiment. Atwo-component mixture of fermionic atoms (red and blue) is pre-pared close to half-filling in a cubic lattice with two different tunnelcoupling configurations. a, Dimerized lattice with the strong dimerlinks td and weaker links t. Low temperatures lead to an excessnumber of singlets over triplets. b, Anisotropic lattice with strongand weak tunneling ts and t along different spatial axes. Antifer-romagnetic spin correlations in the transverse direction are formedalong the strong link direction. In both figures exemplary thermalexcitations in the form of spin excitations or holes are shown.

arX

iv:1

212.

2634

v2 [

cond

-mat

.qua

nt-g

as]

27

Jun

2013

2

a

cb

Merging adjacent sites

Detection Absorptionimages

ps

pt0

mF

-9/2

-7/2

-5/2

-9/2

-7/2

-5/2

Singlet-tripletoscillation

Differential gradient

Preparation

t

tdt

t

tdt

Merging alongstrong links

x

y

z

x

E

FreezingDimerization removal

VX,X,Y,Z = [3.7(1), 0.13(1), 9.8(3), 11.0(3)]ER

0.00.10.20.30.4

Dou

ble

occu

panc

yaf

ter

mer

ging

ηD

0 5 10 15 20 25Time (ms)

0.30.40.50.6

Hig

her

band

fract

ionη

HBMerging along

weak links

FIG. 2. Detection scheme. Summary of the technique used for measuring the atomic singlet and triplet fractions ps and pt0 . a, Schematicview of the different detection steps for the exemplary case of two singlet states in a dimerized lattice. Depending on the oscillation time, theabsorption images on the right show either a large double occupancy in the lowest band corresponding to singlets (top row), or an increasedhigher band fraction indicating triplet states (bottom row). b, The two possible merging configurations in a dimerized lattice. Singlets andtriplets are detected on a set of adjacent sites arranged on a chequerboard pattern in the plane. c, Exemplary singlet-triplet oscillation in astrongly dimerized lattice with U/t = 11.0(8) and td/t = 22(2). We observe an oscillation in the double occupancy after merging, ηD, andin the higher band fraction, ηHB, when merging along the strong links (blue data), whereas no oscillations are visible for the weak links (ochredata). The phase of the oscillation is shifted owing to the double occupancy removal procedure19. The red dashed lines denote the extractedsinglet and triplet fraction ps and pt0 . Error bars show the standard deviation of at least five measurements.

The experiment is performed with a harmonically confined,balanced two-component mixture of a quantum degenerateFermi gas of 40K. The atoms are prepared in two magneticsublevels mF = −9/2,−7/2 of the F = 9/2 hyperfine man-ifold, denoted by ↑ and ↓, at temperatures below 10% of theFermi temperature. We load 50, 000 − 100, 000 repulsivelyinteracting atoms at an s-wave scattering length of 106(1) a0into the three-dimensional optical lattice, where a0 denotesthe Bohr radius. The lattice is created by a combination of in-terfering laser beams operating at 1064 nm with lattice depthsVX, VX, VY and VZ18,19. We independently control the tun-neling strengths along all three spatial axes. In addition wecan introduce a checkerboard dimerization in the xy plane bystrengthening every second tunneling link along the x axis,see Fig. 1a. The checkerboard pattern replicates along thez axis. Our system is well described by a three-dimensionalsingle-band Hubbard model with repulsive on-site interactionenergy U , unless explicitly stated. The tunneling along theweak links in both lattice geometries is set to t/h = 67(3) Hzfor all measurements, where h denotes Planck’s constant.

As shown in Fig. 2a and b, the fraction of atoms formingsinglets and triplets on neighboring lattice sites (ps and pt0 )is detected by transforming the lattice to a checkerboard ge-ometry, similar to a previously developed technique20. In thedimerized lattice our detection scheme locally projects ontothe two-site eigenstates of isolated dimers, whereas for theanisotropic lattice the singlet state is directly projected onto(|↑, ↓〉 − |↓, ↑〉)/

√2 (Ref. 19). In the first detection step, the

atomic motion in the initial lattice is frozen by rapid con-version to a simple cubic structure with negligible tunneling.Next, all atoms on doubly occupied sites are removed. Wethen apply a magnetic field gradient, which creates a differen-tial bias energy ∆ for atoms of opposite spins on adjacent sitesand causes coherent oscillations between the singlet |s〉 =

(|↑, ↓〉− |↓, ↑〉)/√

2 and the triplet |t0〉 = (|↑, ↓〉+ |↓, ↑〉)/√

2state at a frequency ν = ∆/h. If the initial amount of singletsand triplets is equal, no overall oscillation will be visible, as|s〉 and |t0〉 oscillate in antiphase.

After a certain oscillation time, we remove the gradient andmerge two adjacent sites. Owing to the symmetry of the two-

3

particle wavefunction, the singlet state on neighboring sitesevolves to a doubly occupied site with both atoms in the low-est band, while the triplet state transforms into a state withone atom in the lowest and one atom in the first excited band.The fraction of atoms forming double occupancies in the low-est band of the merged lattice,ηD, is detected by a radiofre-quency transfer to the previously unpopulated mF = −5/2spin state9. The fraction of atoms in the higher band ηHB isobtained from a band mapping technique7. For the final read-out we take absorption images after Stern-Gerlach separationof the spin states during ballistic expansion. For an imbalancebetween the initial singlet and triplet populations, ηD and ηHB

will show oscillations with opposite amplitudes. As the dou-ble occupancy in the lowest band contains only contributionsfrom two particles with opposite spins, we can infer the frac-tion of atoms forming singlets and triplets from the maximaand minima of a sinusoidal fit to ηD. The higher band frac-tion has an additional offset caused by dimers containing twoatoms with the same spin or one atom with an antisymmetricspatial wavefunction.

When loading atoms into a strongly dimerized lattice andmerging along the strong links, we observe oscillations in ηDand ηHB, see Fig. 2c. This reveals an excess number of sin-glets, corresponding to magnetic order on neighboring sites.We quantify this order by the normalized imbalance

I =ps − pt0ps + pt0

. (1)

The order in the strongly dimerized lattice originates fromtemperatures below the intra-dimer exchange energy Jd =−U/2 +

√16t2d + U2/2, which denotes the singlet-triplet

splitting on a single dimer. While such temperatures re-quire very low entropies for isotropic lattices21, in our systemthe access to the regime of magnetic ordering is facilitatedby the presence of the weaker inter-dimer exchange energyJ Jd. This leads to an entropy redistribution from stateson the strong to the weak links and gives access to the tem-perature regime J < kBT < Jd for experimentally attainableentropies (here kB denotes the Boltzmann constant). As ex-pected for strong dimerization, we find no visible oscillationswhen merging along the weak links, which indicates the ab-sence of magnetic correlations on these links, see Fig. 2c. Theobserved constant values of ηD = 0.07(1) and ηHB = 0.63(3)are consistent with a state where nearly all singlets are locatedon neighboring dimer links, with vanishing correlations be-tween them.

To analyze the effect of temperature on the magnetic corre-lations, we measure the dependence of the singlet-triplet im-balance on entropy, see Fig. 3a. The imbalance I and the ab-solute singlet fraction ps reduce for larger entropies, as tripletstates become thermally populated. The singlet fraction is ad-ditionally diminished by a shrinking half-filled region in thetrapped system22. We find good agreement with a secondorder high-temperature series expansion of coupled dimerswhen including an entropy increase of ∆s = 0.4 kB with re-spect to the initial entropy in the harmonic trap, sin. Thisheating is associated to the lattice loading23 and is larger forthe lowest entropies, consistent with previous results22. From

100 101

Tunneling ratio td/t

0.0

0.2

0.4

0.6

0.8

1.0

Sin

glet-tr

iple

tim

bala

nceI

100 101 102

Exchange energy ratio Jd/J

0.5 1.0 1.5 2.0 2.5 3.0 3.5Entropy per particle sin (kB)

0.6

0.7

0.8

0.9

1.0

Sin

glet-tr

iple

tim

bala

nceI

I= ps−pt0ps+pt0

0.2 0.4 0.6 0.8 1.0Temperature kBT/Jd

0.5 1.5 2.5sin (kB)

0.150.200.250.30

Sin

glet

fract

ion

p s

ttdt

ttdt

a

b

Δs = 0.0 kB

Δs = 0.4 kB

Δs = 0.8 kB

Δs = 1.2 kB

FIG. 3. Dimerized simple cubic lattice. a, Singlet-triplet imbal-ance on the strong dimer links vs. initial entropy before loading intothe lattice sin and temperature kBT/Jd in a dimerized lattice withU/t = 11.0(8) and td/t = 22(2). The imbalance and the absolutesinglet fraction (inset) decrease with increasing entropy. Solid linesare the prediction of a high-temperature series expansion taking intoaccount different amounts of added entropy ∆s during the latticeloading. b, Imbalance versus dimerization td/t and Jd/J , showingan increase for strongly dimerized simple cubic lattices. The solidline is the theory prediction for an entropy per particle of 1.8 kB inthe lattice, which includes the heating during loading. Vertical errorbars denote the fit error from singlet-triplet oscillations consistingof 63 measurements, the errors in td/t stem from lattice calibrationuncertainties and the errors in sin are the standard deviation of fivemeasurements. For individual curves of ps and pt0 for all measure-ments see19.

the measured imbalances we infer temperatures below 0.4Jd.For reduced dimerizations the coupling between dimers

leads to increased inter-dimer correlations. The excitation en-ergy of triplets is then lowered as they delocalize over the lat-tice, thus changing the nature of magnetic ordering. In Fig.3b we use the tunable lattice to investigate the dependence ofthe imbalance I on the tunneling ratio td/t. As the dimeriza-tion is progressively removed the imbalance decreases in goodagreement with theory and eventually falls below our experi-mental resolution. This decrease can be attributed to the inter-

4

dimer exchange energy Jd becoming smaller than the temper-ature T . For vanishing temperatures the system is expectedto undergo a quantum phase transition from a gapped spin-liquid state to a long-range ordered antiferromagnet as td/t isreduced below a critical value, where the spin gap becomeszero2,24.

1 2 3 4 5 6 7 8Tunneling ratio ts/t

0.00

0.01

0.02

0.03

0.04

Sx iS

x i+1

Sy iS

y i+1

Sxi Sx

i+1 Syi Sy

i+1 =ps−pt0

2

3.05.07.011.0VX (ER)

0 2 4 6 8ts/t

0.00.10.20.3

Nor

mal

ized

spin

corr

elat

orS

1.0 1.5 2.0 2.5Entropy per particle sin (kB)

0.00

0.01

0.02

0.03

0.04

0.05

Sx iS

x i+1

Sy iS

y i+1

ttsts

ti i+1 i i+1

tsts

a

b

FIG. 4. Nearest-neighbor antiferromagnetic order. a, Trans-verse spin correlator versus tunneling ratio ts/t and lattice depthVX in a three-dimensional anisotropic simple cubic lattice withVY,Z = 11.0(3)ER. Positive values correspond to antiferromag-netic ordering. The inset shows the normalized spin correlator S,denoting the fraction of antiferromagnetic ordering at the relevantdensity. Here U/t decreases from 16(1) to 10.5(8). Solid linesare the prediction of a high-temperature series expansion for an en-tropy per particle of 1.8 kB, as used in Fig. 3b, and are shown up tots/kBT = 1/2. b, Transverse spin correlator versus entropy beforeloading into the lattice atU/t = 10.5(8) and ts/t = 7.3(6), togetherwith a schematic view of the spin ordering. Error bars as in Fig. 3.

The key to the observation of quantum magnetism in oursystem is the presence of two different exchange energyscales. Without dimerization, this situation also occurs foranisotropic simple cubic lattices with tunneling t along twoaxes and a stronger tunneling ts along the third direction. In

this case the symmetry between neighboring links is restoredand the detected singlet and triplet fractions are the same forboth merging configurations. We observe a clear populationdifference (ps − pt0)/2 after loading a gas with entropies sinbelow 1.0 kB into an anisotropic lattice, which increases to4% for larger tunneling ratios ts/t, see Fig. 4a. The popu-lation difference is equal to the transverse spin correlator be-tween neighboring sites i and i+ 1 along the strong tunnelingdirection

− 〈Sxi Sxi+1〉 − 〈Syi S

yi+1〉 = (ps − pt0) /2. (2)

This quantity hence directly characterizes the fraction ofatoms with antiferromagnetic ordering on neighboring sites inthe entire atomic cloud. Our observations also extend to weaklattices, where correction terms to the single-band Hubbardmodel become relevant25. In this regime a variety of magneticphases have been predicted26,27.

The results can be compared to a second order high-temperature series expansion19. We find good agreement inthe regime of small anisotropies. For larger anisotropies theexpansion breaks down, as the strong tunneling and the tem-perature become comparable. In this regime we expect thetemperature to lie between the large and small exchange scalesJ < kBT < Js. The system then behaves as an array of one-dimensional spin-ordered chains without correlations betweenthem28, where the majority of the entropy is stored in con-figurations involving the weak links. Low-dimensional sys-tems have been predicted to show enhanced nearest-neighbourcorrelations29.

For temperatures much larger than the strong exchange en-ergy the magnetic correlations should disappear. In Fig. 4bwe study the dependence on the initial entropy sin and findthe correlations to vanish above 2.5 kB, where kBT Js isexpected.

Owing to the presence of the harmonic trap, most spin cor-related atoms are located in the center, where the filling isclose to one particle per site. The density-normalized frac-tion of antiferromagnetic ordering is obtained when dividingby the fraction of atoms with two particles of arbitrary spinon adjacent sites. Under the assumption that all spin correla-tors 〈Sx,y,zi Sx,y,zi+1 〉 are equal – which applies if all symmetrybreaking fields are much smaller than all other energy scales –the normalized spin correlator S can be directly obtained fromthe measurement of singlets and triplets (here nsi is one for asingle particle of any spin on site i and zero otherwise)

S =−4〈Szi Szi+1〉〈nsinsi+1〉

=ps − pt0ps + 3pt0

. (3)

The normalized antiferromagnetic correlations along thestrong tunneling direction reach 25%, see inset Fig. 4a. Thiscorresponds to approximately 5, 000 ordered atoms.

In this work, we have demonstrated the observation ofshort-range quantum magnetism of repulsively interacting ul-tracold fermions in cubic lattices and investigated the depen-dence on temperature, lattice dimerization and anisotropy.Our approach is based on a local entropy redistributionscheme within the lattice structure and can be generalized to

5

access the low temperature regime in different geometries,for example two-dimensional systems. The tunable geome-try optical lattice allows the extension of our studies to spin-ladder systems, dimerized one-dimensional chains and zig-zag chains, where the interplay between quantum fluctuationsand magnetic ordering plays a particularly important role28,30.At even lower temperatures, the existence of spin-liquids in

honeycomb or triangular lattices could be investigated31.Acknowledgments We would like to thank Nils Blumer,

Thierry Giamarchi, Corinna Kollath, Henning Moritz, Chris-tian Ruegg, Manfred Sigrist, Niels Strohmaier and ShunUchino for insightful discussions and Ulf Bissbort for helpwith the calculation of the Hubbard parameters in the dimer-ized lattice. We acknowledge SNF, NCCR-QSIT, NCCR-MaNEP, and SQMS (ERC advanced grant) for funding.

∗ Present Address: ICFO-Institut de Ciencies Fotoniques, Mediter-ranean Technology Park, 08860 Castelldefels (Barcelona), Spain

[email protected] A. Auerbach, Interacting Electrons and Quantum Magnetism

(Springer, 1994).2 S. Sachdev, Quantum magnetism and criticality, Nature Physics

4, 173 (2008).3 P. W. Anderson, et al., The physics behind high-temperature su-

perconducting cuprates: the plain vanilla version of RVB, Journalof Physics: Condensed Matter 16, R755 (2004).

4 L. Balents, Spin liquids in frustrated magnets, Nature 464, 199(2010).

5 M. Lewenstein, et al., Ultracold atomic gases in optical lattices:mimicking condensed matter physics and beyond, Advances inPhysics 56, 243 (2007).

6 I. Bloch, J. Dalibard, W. Zwerger, Many-body physics with ultra-cold gases, Reviews of Modern Physics 80, 885 (2008).

7 T. Esslinger, Fermi-Hubbard physics with atoms in an optical lat-tice, Annual Review of Condensed Matter Physics 1, 129 (2010).

8 M. Kohl, H. Moritz, T. Stoferle, K. Gunter, T. Esslinger,Fermionic atoms in a three dimensional optical lattice: observ-ing Fermi surfaces, dynamics, and interactions, Physical ReviewLetters 94, 080403 (2005).

9 R. Jordens, N. Strohmaier, K. Gunter, H. Moritz, T. Esslinger, AMott insulator of fermionic atoms in an optical lattice, Nature 455,204 (2008).

10 U. Schneider, et al., Metallic and insulating phases of repulsivelyinteracting fermions in a 3D optical lattice, Science 322, 1520(2008).

11 J. Simon, et al., Quantum simulation of antiferromagnetic spinchains in an optical lattice, Nature 472, 307 (2011).

12 J. Struck, et al., Quantum simulation of frustrated classical mag-netism in triangular optical lattices, Science 333, 996 (2011).

13 S. Trotzky, et al., Time-resolved observation and control of su-perexchange interactions with ultracold atoms in optical lattices,Science 319, 295 (2008).

14 S. Nascimbene, et al., Experimental realization of plaquette res-onating valence-bond states with ultracold atoms in optical super-lattices, Physical Review Letters 108, 205301 (2012).

15 T.-L. Ho, Q. Zhou, Squeezing out the entropy of fermions in opti-cal lattices, Proceedings of the National Academy of Sciences 106,6916 (2009).

16 J.-S. Bernier, et al., Cooling fermionic atoms in optical lattices byshaping the confinement, Physical Review A 79, 061601 (2009).

17 H. T. Diep, G. Misguich, C. Lhuillier, Frustrated Spin Systems(World Scientific, 2005), pp. 229–306.

18 L. Tarruell, D. Greif, T. Uehlinger, G. Jotzu, T. Esslinger, Cre-ating, moving and merging Dirac points with a Fermi gas in atunable honeycomb lattice, Nature 483, 302 (2012).

19 See Supplementary Materials.20 S. Trotzky, Y.-A. Chen, U. Schnorrberger, P. Cheinet, I. Bloch,

Controlling and detecting spin correlations of ultracold atoms inoptical lattices, Physical Review Letters 105, 265303 (2010).

21 S. Fuchs, et al., Thermodynamics of the 3D Hubbard model onapproaching the Neel transition, Physical Review Letters 106,030401 (2011).

22 D. Greif, L. Tarruell, T. Uehlinger, R. Jordens, T. Esslinger, Prob-ing nearest-neighbor correlations of ultracold fermions in an opti-cal lattice, Physical Review Letters 106, 145302 (2011).

23 R. Jordens, et al., Quantitative determination of temperature inthe approach to magnetic order of ultracold fermions in an opticallattice, Physical Review Letters 104, 180401 (2010).

24 C. Ruegg, et al., Pressure-induced quantum phase transition in thespin-liquid TlCuCl3, Physical Review Letters 93, 257201 (2004).

25 F. Werner, O. Parcollet, A. Georges, S. R. Hassan, Interaction-induced adiabatic cooling and antiferromagnetism of coldfermions in optical lattices, Physical Review Letters 95, 056401(2005).

26 C. J. M. Mathy, D. A. Huse, Accessing the Neel phase of ultracoldfermionic atoms in a simple-cubic optical lattice, Physical ReviewA 79, 063412 (2009).

27 P. N. Ma, S. Pilati, M. Troyer, X. Dai, Density functional theoryfor atomic Fermi gases, Nature Physics 8, 601 (2012).

28 T. Giamarchi, Quantum Physics in One Dimension (ClarendonPress, 2003).

29 E. V. Gorelik, et al., Universal probes for antiferromagnetic corre-lations and entropy in cold fermions on optical lattices, PhysicalReview A 85, 061602 (2012).

30 P.-B. He, Q. Sun, P. Li, S.-Q. Shen, W. M. Liu, Magnetic quan-tum phase transition of cold atoms in an optical lattice, PhysicalReview A 76, 043618 (2007).

31 Z. Y. Meng, T. C. Lang, S. Wessel, F. F. Assaad, A. Mura-matsu, Quantum spin liquid emerging in two-dimensional corre-lated Dirac fermions, Nature 464, 847 (2010).

32 J. S. Krauser, et al., Coherent multi-flavour spin dynamics in afermionic quantum gas, Nature Physics 8, 813 (2012).

33 J. A. Henderson, J. Oitmaa, M. C. B. Ashley, High-temperatureexpansion for the single-band Hubbard model, Physical Review B46, 6328 (1992).

34 D. F. B. ten Haaf, J. M. J. van Leeuwen, High-temperature seriesexpansions for the Hubbard model, Physical Review B 46, 6313(1992).

6

SUPPLEMENTARY MATERIALS

I. EXPERIMENTAL DETAILS

A. Preparation

After sympathetic cooling with 87Rb, 2 × 106 fermionic40K atoms are transferred into an optical dipole trap operatingat a wavelength of 826 nm. A balanced incoherent spin mix-ture of atoms in the Zeeman levels mF = −9/2 and −7/2 ofthe F = 9/2 hyperfine manifold is then prepared and evapo-ratively cooled9. When taking data as a function of entropy,the gas is heated through inelastic losses by setting the mag-netic bias field to a value close to the Feshbach resonance at202.1 G. We measure the entropy per particle in the dipoletrap sin using Fermi fits to the momentum distribution of thecloud after expansion. The field is finally increased to 221.4 Gresulting in a scattering length of 106(1) a0. The optical lat-tice is subsequently turned on in 200 ms using a spline shapedlaser-intensity ramp.

B. Trapping potential

The red-detuned optical lattice is created by four retro-reflected laser beams at λ = 1064 nm18. This gives rise toa potential of the form

V (x, y, z) = −VX cos2(kx+ θ/2)− VX cos2(kx)

−VY cos2(ky)

−2α√VXVY cos(kx) cos(ky) cosϕ

−VZ cos2(kz), (S1)

where VX, VX, VY and VZ denote single-beam lattice depths(as calibrated using Raman-Nath diffraction on a 87Rb Bose-Einstein condensate), k = 2π/λ and the measured visibilityof the interference pattern α is 0.90(5). The phase ϕ is sta-bilized to 0.00(3)π, whilst θ is set to 1.000(1)π. Gravitationpoints along the y direction.

The laser beams create an overall harmonic trapping po-tential which scales with the lattice depths according to theapproximate expressions

ωx ∝√VY + 1.11VZ

ωy ∝√VX + 0.81(VXVY/VX) + 0.78VZ

ωz ∝√VX + 0.81(VXVY/VX) + 1.24VY. (S2)

For VX,X,Y,Z = [3.7(1), 0.13(1), 9.8(3), 11.0(3)]ER, asin Fig. 3a, the lattice contributes trapping frequen-cies of ωLattice

x,y,z /2π = [62.7(9), 57(1), 54.3(3)] Hz. Ad-ditionally, the optical dipole trap creates a trapping ofωDipolex,y,z /2π = [30.7(2), 105.9(3), 34.6(2)] Hz, in all data sets

shown, except for Fig. 2c and 4a, where ωDipolex,y,z /2π =

[28.1(2), 90.1(3), 31.6(2)] Hz.

C. Detection lattice ramp

For measurements in the dimerized lattice, the lattice isramped up in two steps. All beam intensities are linearlyincreased over the course of 0.5 ms up to the point whereVY = 30(1)ER, VZ = 40(1)ER and all other intensitiesin the xy plane are ramped such that the potential is not de-formed. In a second linear ramp lasting 10 ms, the latticeis changed to a simple cubic geometry where VX,X,Y,Z =

[25(1), 0, 30(1), 40(1)]ER.This ramp can be considered sudden for the inter-dimer

links but adiabatic for the intra-dimer links. Our observablehence locally projects onto the two-site eigenstates of individ-ual dimers, which includes an admixture of double occupan-cies. We use an exact calculation of a two-site Hubbard modelto estimate the adiabaticity of the ramp. The unnormalizedsinglet ground state of this system is given by

4td

−U +√

16t2d + U2(|↑, ↓〉 − |↓, ↑〉) + (|↑↓, 0〉+ |0, ↑↓〉) .

(S3)

There is hence a significant contribution of double occupancyin the regime where U ∼ td, as applies for the most stronglydimerized lattices investigated in this paper, whilst the con-tribution vanishes for the deep simple cubic lattice used fordetection where U/t ≈ 600. For the given ramp-times andincluding site-offsets due to the harmonic trapping potential,the probability of populating excited states on a dimer duringsuch a ramp remains below 0.1% for all values of U and tdexplored in the dimerized systems.

In the anisotropic lattice, we directly ramp to VX,Y,Z =

[25(1), 30(1), 40(1)]ER in 0.5 ms (VX = 0ER throughout).In contrast to the dimer lattice ramp, this process can be con-sidered sudden for all links, as the duration of the ramp isalways well below the tunneling time in the initial lattice. Ourdetection method then corresponds to locally projecting thewavefunction of the system onto (|↑, ↓〉− |↓, ↑〉)/

√2 on pairs

of sites when measuring ps, hence excluding any contributionsfrom double occupancies. The probability of this projection,as calculated from the two-site Hubbard model, lies above80% for all shown data and is higher for deeper lattices. Forboth lattice geometries, the triplet state remains unaffected bychanges in U and t.

D. Singlet-triplet oscillations

For both systems, once the ramp to the deep simple cubiclattice has been completed, double occupancies are removedvia spin-changing collisions, which occur after transferringatoms from the mF = −7/2 to the mF = −3/2 state32.We verify that this procedure removes all double occupan-cies but leaves singly occupied sites unaffected by measur-ing the number of double occupancies and the total numberof atoms before and after applying the removal sequence. Amagnetic field gradient causing spin-dependent energy offsets

7

of ∆−9/2,−7/2/h = [1291(1), 1156(1)] Hz between neigh-boring sites is then applied, giving rise to coherent oscillationsbetween singlets and triplets20. Subsequently, pairs of adja-cent lattice sites are adiabatically merged into one by linearlyreducing VX to zero whilst increasing VX to 25(1)ER. Thedouble occupancy in the merged lattice is then measured as9,but taking into account an independently calibrated detectionefficiency of 89(2)%. We verify that merged sites contain-ing two atoms of opposite spin but in different bands are notdetected as double occupancies by artificially creating a statecontaining large amounts of triplets but no singlets and mea-suring ηD.

We apply a sinusoidal fit to the double occupancy wherethe frequency and phase are fixed and take into account thedamping of the oscillations, which was calibrated indepen-dently and is included by multiplying the amplitude by 1.16.A phase shift arises owing to a weak residual magnetic fieldgradient present during the double occupancy removal pro-cedure, whereas the contribution from switching the singlet-triplet oscillation gradient on and off is negligible. We confirmthat the maximum of the oscillation corresponds to its startingpoint (and hence to the number of singlets) by merging thelattice immediately after it has been ramped to a deep simplecubic structure.

E. Theoretical model

For sufficiently deep lattices the system is well describedby a single band Fermi-Hubbard model, which is given by

H = −ti,j∑〈ij〉,σ

(c†iσ cjσ + h.c.) + U∑i

ni↑ni↓ (S4)

for a homogeneous system, using standard notation.The in-teraction energy U and nearest-neighbor tunneling ti,j areevaluated from Wannier function integrals5. The dimerizedlattice theory curves are calculated from a perturbative cou-pling of isolated dimer links with tunneling t. The result-ing partition function expansion is then calculated up to sec-ond order in t/kBT . While the dimer link contains 16 statesin total and the tunneling operator is non-diagonal betweenneighboring dimers, the evaluation of the relevant matrix el-ements is directly analogous to the case of single sites33.The harmonic trap is included in a local density approxima-tion, which leads to a quadratically varying chemical poten-tial. All thermodynamic quantities are obtained after integra-tion over the entire trap using independent calibrations of theatom number, trap frequencies and lattice depths23. In thelimit of very strong dimerization, theory predicts pt0/ps ∝exp(−Jd/kBT ), which can be used for lattice thermometry.

For the anisotropic lattice we evaluate the correlators〈Sz

i Szi+1〉 and S in a second order series expansion of coupled

single sites34. The thermodynamic observables are obtainedin a similar way as previously described23, using the averagetunneling

√(t2s + 2t2)/3. Eq. (2) in the main text is computed

by evaluating the matrix elements of the spin operators.

II. DIMERIZED CUBIC LATTICE - SERIES EXPANSION

We start with the definition of the homogeneous single-band Hubbard Hamiltonian in a dimerized cubic lattice withstrong and weak links between nearest neighbors 〈i, j〉 and〈i, j〉 respectively, see Fig. S1,

HD = Ho + Hc

Ho = −td∑

σ,〈i,j〉(c†i,σ cj,σ + h.c.)

+U∑i

ni↑ni↓ − µ∑i

(ni↑ + ni↓) (S5)

Hc = −t∑

σ,〈i,j〉(c†i,σ cj,σ + h.c.).

We have split the Hamiltonian into the dimer part Ho andthe coupling between dimers Hc. The on-site interaction en-ergy is given by U , the tunneling matrix elements betweennearest neighbors by t and td and the chemical potential isparametrized with µ. The fermionic creation operator for anatom on the lattice site i is given by c†iσ , where σ ∈ ↑, ↓denotes the magnetic sublevel and h.c. is the Hermitian con-jugate. The particle number operator is ni = ni↑+ ni↓, niσ =

c†iσ ciσ . Denoting the inverse temperature with β = 1/kBT ,the thermal average of an observableO then reads in the grandcanonical potential

a b

Dimerized cubic lattice Anisotropic cubic lattice

t

t

t

t t

td ts ts

xy

z

FIG. S1. Lattice geometries. Overview of the three-dimensionallattice geometries explored in the experiment. a, In the dimerizedcubic lattice the tunneling between neighboring sites is increased totd on a checkerboard pattern in the xy plane as compared to theweaker tunneling t. The pattern replicates along the z axis. b, In theanisotropic cubic lattice the tunneling is increased to ts along the xaxis, while remaining t in the other two directions.

〈O〉 =TrOe−βHDTre−βHD

. (S6)

We now treat the coupling Hamiltonian Hc = tT as a pertur-bation, which leads to an expansion of the above expressionin powers of the dimensionless parameter βt33. The expan-sion is expected to be close to the exact result in the regimet kBT td, U . For the partition function up to secondorder (denominator in Eq. S6) we find

Z = Z0 + (βt)2Z0

β2

∫ β

0

∫ τ1

0

dτ2dτ1〈T ′(τ1)T ′(τ2)〉0. (S7)

8

The expression for the numerator is analogous. The partitionfunction of the unperturbed Hamiltonian is denoted by Z0,whereas 〈...〉0 denotes the thermal average of the unperturbedHamiltonian

〈T ′(τ1)T ′(τ2)〉0 = Trexp(−βHo)T′(τ1)T ′(τ2)/Z0

T ′(τ) = eτHo T e−τHo . (S8)

III. DIMERIZED CUBIC LATTICE - OBSERVABLES

As the expansion is up to second order in the tunnel cou-pling, it is sufficient to evaluate all expressions in a two-dimerbasis. Denoting the single dimer Hamiltonian in the grandcanonical potential with Hs

o, the unperturbed partition func-tion then reads

Z0 =(

Tre−βHso)2. (S9)

The evaluation of the second order terms is done in a doubledimer basis |Ψ1

i ,Ψ2j 〉, where |Ψ1

i 〉 and |Ψ2j 〉 each denote one

of the 16 possible eigenvectors of the first and second dimerlink. This essentially leaves the evaluation of matrix elementsof the following kind

〈Ψ2j ,Ψ

1i |T ′(τ1)T ′(τ2)|Ψ1

i ,Ψ2j 〉 and

〈Ψ2j ,Ψ

1i |OT ′(τ1)T ′(τ2)|Ψ1

i ,Ψ2j 〉, (S10)

which can be computed either directly or numerically. For thesinglet and triplet fraction the observable O takes the form ofa projector for the 16 possible states on a dimer. The entropyand particle number per dimer are evaluated from the grandcanonical potential Ωd = −kBT logZd of a single dimer.Fig. S2 shows the dimer singlet probability and the entropyper site versus filling calculated in second order for differenttemperatures and dimerizations. A comparison between thepredictions of lowest order (atomic limit) and second order isshown in Fig. S3.

IV. ANISOTROPIC CUBIC LATTICE

Similar to the case of the dimerized lattice, we split theHamiltonian for the homogeneous anisotropic cubic latticeinto two parts

HA = HU + Ht

HU = U∑i

ni↑ni↓ − µ∑i

(ni↑ + ni↓) (S11)

Ht = −ts∑

σ,〈i,j〉(c†i,σ cj,σ + h.c.)− t

∑σ,〈i,j〉

(c†i,σ cj,σ + h.c.).

Notations are analogous to the previous section, see Fig. S1.The strong tunneling between nearest neighbors along the xdirection is denoted with ts, whereas the weaker tunnelingalong the other two axes is given by t. We treat the tunnel-ing Hamiltonian Ht as a perturbation to the unperturbed part

HU , which leads to an expansion of the partition function as inEq. S7 in powers of βts and βt. Density and entropy per siteare then obtained from derivatives of the second order grandpotential Ω,

βΩ = −logZ1−2(βts)

2 + 4(βt)2

(Z1)2

(ζ + ζ3w + 2ζ2

1− wβU

).

(S12)Here Z1 is the unperturbed single site partition function, ζ =exp(βµ) the fugacity and w = exp(−βU).

The evaluation of the two correlators defined in the maintext 〈Szi Szi+1〉 and S is slightly more complicated, as it in-volves two neighboring sites. However, the coefficients forthese correlators have already been computed34.

〈Szi Szi+1〉 = − ζ2

(Zs0)2

(1

βU+w − 1

(βU)2

)(βts)

2 (S13)

S =

(1

βU+w − 1

(βU)2

)(βts)

2. (S14)

V. HARMONIC TRAP

The effect of the harmonic trap is included in a local densityapproximation with a quadratically varying chemical potential

µ(r) = µ0 −1

2mω2(

λ

2)2r2, (S15)

where ω is the geometric mean of the trapping frequencies,µ0 the chemical potential in the center of the trap and r thenormalized distance of a site to the trap center. Any trap av-eraged observable Otrap is then obtained from integration ofthe contributions per site Ohom(µ)

Otrap =

∫ ∞0

4πr2Ohom(µ(r))dr. (S16)

Owing to the harmonic trap, the energy offset between neigh-boring sites on the dimer links changes over the cloud size.The relative correction of this effect to all shown quantitieswas computed to be less than a few percent.

VI. SINGLET AND TRIPLET FRACTIONS

The fraction of atoms forming singlets and triplets ps andpt0 are obtained from an integration over the left sites of eachmerged pair, which are part of the set A (see Fig. 2b),

ps = 2∑i∈A

⟨P si

⟩/N P s

i = |Ψsi〉 〈Ψs

i |

pt0 = 2∑i∈A

⟨P t0i

⟩/N P t0

i =∣∣Ψt0

i

⟩ ⟨Ψt0i

∣∣ . (S17)

Here P si and P t0

i are the projection operators on the singletand triplet states |Ψs

i〉 and∣∣Ψt0

i

⟩on neighboring sites i and

i + 1, 〈...〉 denotes the thermal average and N the total atom

9

td/t = 10

0.0 0.5 1.0 1.5 2.0Particles per site

0.0

0.2

0.4

0.6

0.8

1.0

Dim

ersi

ngle

tpr

obab

ility

td/t = 2

0.0 0.5 1.0 1.5 2.0Particles per site

0.0

0.2

0.4

0.6

0.8

1.0

Dim

ersi

ngle

tpr

obab

ility

td/t = 10

0.0 0.5 1.0 1.5 2.0Particles per site

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Ent

ropy

per

site

(kB)

td/t = 2

0.0 0.5 1.0 1.5 2.0Particles per site

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Ent

ropy

per

site

(kB)

kBT/t = 1

kBT/t = 3

kBT/t = 10

FIG. S2. Dimerized lattice. High-temperature series predictions up to second order for the homogeneous dimerized cubic lattice. Thedependence on filling of the dimer singlet probability and of the entropy per site is shown. We set U/t = 5 and td/t = 10 or 2 and usedifferent temperatures kBT/t. The entropy at half filling for large dimerization is strongly reduced.

2 4 6 8 10 12 140.0

0.2

0.4

0.6

0.8

1.0

Dim

ersi

ngle

tpr

obab

ility

1 2 3 4 5 6 7 8 9 100.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Ent

ropy

per

site

td/t = 2

td/t = 10

td/t = 2

td/t = 10

kBT/t kBT/t

FIG. S3. Higher order contributions. Comparison of the high-temperature series predictions in lowest and second order (dashed and solidline) for the dimer singlet probability and for the entropy per site in a homogeneous dimerized cubic lattice. The filling is set to one particleper site and the interaction to U/t = 5. The second order contributions are expected to be larger for lower temperatures and lead to a reductionof the singlet probability.

number. For the measurements in the anisotropic simple cubiclattice, the projection operators are related to the spin opera-tors ~Si = 1/2

∑s,s′ c

†i,s~σci,s′ , where ~σ = (σx, σy, σz) are the

Pauli matrices

P si =

nini+1

4− ~Si~Si+1

P t0i = ~Si~Si+1 − 2Szi S

zi+1 +

nini+1

4. (S18)

From this the equality of Eq. (2) in the main text is immedi-ately obtained. In the dimerized lattice the projection operatoron the triplet reads the same, whereas for the singlet the two-site system needs to be diagonalized.

Fig. S4 shows the individual atomic fractions of singletsand triplets (ps and pt0 ) measured in the dimerized lattice,which are used to compute the normalized imbalance I. The

data in Fig. S4a corresponds to the entropy scan of Fig. 3ain the main text, whilst the set of measurements for differ-ent tunneling ratios in Fig. S4b corresponds to Fig. 3b. Weadditionally measured the fraction of atoms on doubly occu-pied sites Ddimer in the lattice after freezing out the atomicmotion but before applying the cleaning procedure and induc-ing singlet-triplet oscillations. Note that for the dimer lattice,this fraction does not contain any contributions from the ad-mixture of double occupancies present in the initial singletstate. For the entropy scan we find a nearly constant value ofDdimer = 0.026(5). For the measurements taken at differ-ent td/t the double occupancy ranges between 0.066(4) and0.29(2).

Fig. S5 shows ps and pt0 for the measurements in theanisotropic cubic lattice, which are used to compute the spincorrelators −〈Sxi Sxi+1〉 − 〈S

yi S

yi+1〉 and S. Fig. S5a cor-

10

0.16

0.20

0.24

0.28

0.32

Sin

glet

fract

ion

p s

0.5 1.0 1.5 2.0 2.5

Entropy per particle sin (kB)

0.00

0.01

0.02

0.03

0.04

0.05

0.10

0.15

0.20

0.25

0.30

Sin

glet

fract

ion

p s

100 101

Tunneling ratio td/t

0.00

0.02

0.04

0.06

0.08

Trip

let

fract

ion

p t0

Trip

let

fract

ion

p t0

a b

FIG. S4. Dimerized simple cubic lattice. a, Singlet and triplet fractions on the strong dimer links vs. initial entropy before loading into thelattice. The data shown here is used to compute the normalized imbalance I displayed in Fig. 3a. b, Singlet and triplet fractions as a functionof dimerization td/t, corresponding to the measurements of Fig. 3b. Vertical error bars denote the fit error from singlet-triplet oscillationsconsisting of 63 measurements, the errors in td/t stem from lattice calibration uncertainties and the errors in sin are the standard deviation offive measurements.

0.06

0.08

0.10

0.12

0.14

Sin

glet

fract

ion

p s

1.0 1.5 2.0 2.5Entropy per particle sin (kB)

0.04

0.05

0.06

0.07

0.08

0.10

0.12

0.14

Sin

glet

fract

ion

p s

0 1 2 3 4 5 6 7 8Tunneling ratio ts/t

0.05

0.06

0.07

0.08

0.09

0.10

Trip

let

fract

ion

p t0

a b

Trip

let

fract

ion

p t0

FIG. S5. Anisotropic simple cubic lattice. a, Singlet and triplet fractions as a function of initial entropy before loading into the lattice,corresponding to the measurements of Fig. 4b. b, Singlet and triplet fractions vs. tunneling anisotropy ts/t, which are used to compute thespin correlators presented in Fig. 4a. Vertical error bars denote the fit error from singlet-triplet oscillations consisting of 63 measurements, theerrors in ts/t stem from lattice calibration uncertainties and the errors in sin are the standard deviation of five measurements.

responds to the measurements versus entropy displayed inFig. 4b, whilst the scan of the tunneling anisotropy shown inFig. S5b corresponds to the data presented in Fig. 4a. Here,we also measured the double occupancy immediately afterfreezing out the atomic motion Danisotropic, which now cor-

responds to a direct projection onto doubly occupied sites.For the entropy scan Danisotropic lies between 0.11(6) and0.21(1), whilst for the scan of ts/t we measure double oc-cupancies between 0.14(2) and 0.19(2).