You are on page 1of 27

LETTERS

PUBLISHED ONLINE: 6 JUNE 2016 | DOI: 10.1038/NPHYS3783

Many-body localization in a quantum simulator


with programmable random disorder
J. Smith1*, A. Lee1, P. Richerme2, B. Neyenhuis1, P. W. Hess1, P. Hauke3,4, M. Heyl3,4,5, D. A. Huse6
and C. Monroe1
When a system thermalizes it loses all memory of its initial faster than the coupling rate to the outside environment (see
conditions. Even within a closed quantum system, subsystems Supplementary Information).
usually thermalize using the rest of the system as a heat bath. Each of the effective spin-1/2 particles is encoded in the
Exceptions to quantum thermalization have been observed, 2
S1/2 |F = 0, mF = 0i and |F = 1, mF = 0i hyperfine ‘clock’ states of
but typically require inherent symmetries1,2 or noninteracting a 171 Yb+ ion, denoted |↓iz and |↑iz , respectively. We confine a
particles in the presence of static disorder3–6 . However, for chain of ten ions in a linear radiofrequency Paul trap and apply
strong interactions and high excitation energy there are cases, optical dipole forces to generate the effective spin–spin coupling14
known as many-body localization (MBL), where disordered of a disordered Ising Hamiltonian:
quantum systems can fail to thermalize7–10 . We experimentally
generate MBL states by applying an Ising Hamiltonian with X BX X Di
HIsing = Ji,j σix σjx + σiz + σiz (1)
long-range interactions and programmable random disorder to 2 2
ten spins initialized far from equilibrium. Using experimental
i<j i i

and numerical methods we observe the essential signatures where σiγ (γ = x, z) are the Pauli matrices acting on the ith spin,
of MBL: initial-state memory retention, Poissonian distributed Ji,j is the coupling strength between spins i and j, B is a uniform
energy level spacings, and evidence of long-time entanglement effective transverse field, Di is a site-dependent disordered potential,
growth. Our platform can be scaled to more spins, where a and h̄ = 1 (see Supplementary Information). After the chain evolves
detailed modelling of MBL becomes impossible. for some time, we collect the state-dependent fluorescence on
It is exceedingly rare in nature for systems to localize, or retain an intensified charge-coupled device camera for site-resolved
local information about their initial conditions at long times. imaging. This, in addition to our ability to perform high-fidelity
In an important counterexample, Anderson demonstrated that rotations, allows measurement of the single-site magnetization
localization can arise due to the presence of disorder, which can hσiγ i(γ = x, y, z) as well as arbitrary spin correlation functions along
destructively scatter propagating waves and prevent transport of any direction.
energy or particles3 . Although this interference effect can be applied Ji,j is a tunable, long-range coupling that falls off approximately
to generic quantum systems, most experimental work has been algebraically as Ji,j ∝ Jmax /|i − j|α (ref. 15), where Jmax is typically
restricted to the narrow parameter regime of low excitation energies 2π (0.5 kHz). Here, we tune α between 0.95 and 1.81, although
and no interparticle interactions4–6 . for most of the data α ≈ 1.13. We directly measure the complete
Whether such localization persists in the more general case of spin–spin coupling matrix (Fig. 1a), demonstrating the long-range
arbitrary excitation energy and non-zero interparticle interactions interactions required to exhibit MBL in this model. Moreover, it has
was theoretically explored by Anderson3 , and more recently by oth- been shown numerically that (1) exhibits MBL for the experimental
ers7–10 . This MBL phase is predicted to emerge for a broad set of in- parameters16 .
teraction ranges and disorder strengths, although the precise phase The site-specific programmable disorder term Di is sampled from
diagram is not well known11 because equilibrium statistical mechan- a uniform random distribution with Di ∈ [−W , W ], where W
ics breaks down in the MBL phase and numerical simulations are characterizes the strength of the disorder. The disorder is generated
limited to ∼20 particles8,9 . Very recent experiments searching for by site-dependent laser-induced Stark shifts (see Supplementary
MBL have measured constrained mass transport and the breakdown Information), which also allow for preparation of the system
of ergodicity in disordered atomic systems with interactions12,13 . into any desired product state. To ensure we observe the general
Here we report the direct observation of MBL in a long- behaviour of the disordered Hamiltonian, we average over 30
range transverse field Ising model with programmable, random distinct random instances of disorder (Fig. 1b,c), which leads to
disorder. This is a non-integrable model that cannot be mapped to a sampling error that is smaller than the features of interest (see
noninteracting particles (a necessary condition for MBL7 ) and we Supplementary Information).
can easily tune the disorder strength and interaction range over a An important signature of the MBL phase is manifested in the
parameter space that exhibits this phenomenon. Our experiment spectral statistics of adjacent energy levels of the Hamiltonian. In the
is effectively a closed quantum system over the timescales of thermalizing phase, the energy levels are given by the eigenvalues of
interest, because the system localizes approximately 60 times a random matrix, a matrix whose elements are given by a random

1
Joint Quantum Institute, University of Maryland, Department of Physics and National Institute of Standards and Technology, College Park,
Maryland 20742, USA. 2 Department of Physics, Indiana University, Bloomington, Indiana 47405, USA. 3 Institute for Quantum Optics and Quantum
Information of the Austrian Academy of Sciences, 6020 Innsbruck, Austria. 4 Institute for Theoretical Physics, University of Innsbruck, 6020 Innsbruck,
Austria. 5 Physik Department, Technische Universität München, 85747 Garching, Germany. 6 Physics Department, Princeton University, Princeton,
New Jersey 08544, USA. *e-mail: htimsjacob@gmail.com

NATURE PHYSICS | VOL 12 | OCTOBER 2016 | www.nature.com/naturephysics 907


© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
LETTERS NATURE PHYSICS DOI: 10.1038/NPHYS3783

a 0.5 b 8

Spin−spin coupling (kHz)


6
0.4
4

Disorder strength (Jmax)


0.3
2
0.2
0

0.1 −2

−4 Ion positions
1 2 10
3 4 8 9 −6
5 5 6 7
Spin 6
num 7 8 9
4 ber
ber 2 3 in num −8
10 1 Sp
c

30

d W = 8Jmax
25 0.3

Fraction of occurrences
Instance of
20
0.2

15
disorder

0.1
10

1 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14


10 ΔE/Jmax
9
7 8
4 5 6 on
iti
2
3 pos
1 Ion

Figure 1 | An interacting spin model with random disorder. a, Directly measured elements of the spin–spin coupling matrix Jij , equation (1) (increasing
interaction strength from blue to red). The long-range interactions decay as Jmax /r1.13 . b, A specific instance of the random disordered field with a
schematic illustration of the long-range interactions. c, Random values of the disordered field for all 30 instances of disorder for several different disorder
strengths and for each ion (red indicates positive values and blue indicates negative values, with values between −0.5 and 0.5). d, Level statistics
calculated from the measured spin–spin coupling matrix in a and applied disorders in c are Poisson-distributed (black line is the expected level spacings for
a Poisson distribution), as predicted for a MBL system.

distribution, due to level repulsion. However, in the MBL phase, this When B  P J , the Hamiltonian is effectively an XY model17,18
level repulsion is greatly suppressed because eigenstates typically and conserves i σiz , because Ising processes that flip spins along
differ by multiple spins flips. As a result, the level spacing between the large field are energetically forbidden. Thus, being in a spin
adjacent energy eigenvalues are Poisson-distributed8,9 . Using our configuration with half of the spins up and half of the spins down
directly measured spin–spin couplings and applied realizations for maximizes the accessible energy states. In addition, the Néel state is
the strongest experimental disorder W = 8Jmax and B = 4Jmax , we never an eigenstate, even for B  J and W  J , because the uniform
calculate the distribution of adjacent energy level splittings and find B field at each site still allows spin exchange in the z-basis.
them to be Poisson-distributed, as expected for a MBL state (Fig. 1d). If a system is thermal, the Eigenstate Thermalization Hypothesis
Before searching for evidence of localization in the system’s (ETH) provides a general framework where observables reach the
time evolution, we first find parameters that cause the measured value predicted by the microcanonical ensemble19–21 . This allows us
state to thermalize in the absence of disorder. We increase the to calculate the expected thermal value of the reduced density matrix
transverse field B and look for conditions that result in the single-site given the Hamiltonian and an initial state (see Supplementary
magnetization along two orthogonal directions approaching and Information). To further establish that the system is thermalizing,
remaining at their thermal equilibrium values (see Supplementary we measure the reduced density matrix for each spin, ρi = Tr{j6=i} ρ,
Information). without applied disorder and B = 4Jmax , as shown in Fig. 2a. In
Figure 2a shows the measured dynamics of hσiz i for B = 4Jmax our experiment, the spins are initially prepared in a product state
and Di = 0 with the spins initialized in the Néel ordered state, with high fidelity. However, at long times, the measured reduced
|↑↓↑↓↑↓↑↓↑↓iz along the z direction. This configuration has an density matrices show that each of the spins is very close to the
energy equivalent to an infinite temperature thermal state, because zero magnetization mixed state, implying the system has locally
the expectation value of the Hamiltonian is zero. At long times, each thermalized.
expectation value hσiz i approaches zero, losing memory of the initial We apply the random disordered potential, Di 6 = 0, with B = 4Jmax
ordering. As the transverse field B is increased, the system appears and observe the emergence of MBL as we increase the strength
to thermalize more quickly and the level statistics approach those of of disorder. Because the many-body eigenstates in the MBL phase
random matrices rather than Poissonians, as expected for a generic are not thermal, transport of energy and spins is suppressed,
thermodynamic system (see Supplementary Information). and ETH fails. Thus, observables will not relax to their thermal

908 NATURE PHYSICS | VOL 12 | OCTOBER 2016 | www.nature.com/naturephysics

© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ


NATURE PHYSICS DOI: 10.1038/NPHYS3783 LETTERS
a Jmaxt = 0 No disorder Jmaxt ≥ 5 f Jmaxt = 0 Strongest disorder Jmaxt ≥ 5
0.97 Ion 1 ( ) 1.0 1.0
W = 0Jmax Ion 1 ( ) 0.96 Ion 1 ( ) W = 8Jmax Ion 1 ( )

Z magnetization, 〈σ iz 〉

Z magnetization, 〈σ iz 〉
0.78
0.53 0.47
0.03 0.5 0.5 0.22
0.04
0.02 〈↓| 0.01 0.05 0.08 〈↓|
| ↑〉 |↑〉 〈↓| |↑〉 〈↓| |↑ 〉
| ↓〉 〈↑| 0.0 |↓〉 〈↑| |↓〉 〈↑| 0.0 |↓〉 〈↑|

Ion 10 ( ) 1.00 −0.5 Ion 10 ( ) Ion 10 ( ) 0.99 −0.5 Ion 10 ( )


0.78
0.50 0.50
0.00 0.01 0.22
−1.0 −1.0
0.07 〈↓| 0 2 4 6 8 10 0.01 〈↓| 0.05 〈↓| 0 2 4 6 8 10 0.07 〈↓|
| ↑〉 |↑〉 |↑〉 |↑〉
|↓〉 〈↑| Jmaxt |↓〉 〈↑| |↓〉 〈↑| Jmaxt |↓〉 〈↑|

Increasing disorder
b c d e
1.0
W = 1Jmax W = 2Jmax W = 4Jmax W = 6Jmax
Z magnetization, 〈σ iz 〉

0.5

0.0

−0.5

−1.0
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Jmaxt Jmaxt Jmaxt Jmaxt

Figure 2 | Emergence of a MBL state. a, Time-evolved single-site magnetizations hσiz i (different colours represent different ions) for the Hamiltonian in
equation (1) for B = 4Jmax with no applied disorder (Di = 0). The initial-state reduced density matrices for ions 1 and 10 show the spins start in a product
state along the z direction. The time-averaged reduced density matrices for Jmax t > 5 (colours from blue to red indicate increasing values of the elements of
the density matrix) agree with the values predicted by the ETH, implying the system has thermalized locally. b–e, As the disorder strength increases, the
spins retain more information about their initial state, indicating a transition towards MBL. f, Dynamics of hσiz i for the strongest applied disorder,
W = 8Jmax . The initial-state and steady-state time-averaged reduced density matrices for ions 1 and 10 now show that information is preserved about the
initial spin configuration at the end of the evolution. Statistical error bars (1 s.d.) are smaller than the data points.

values9 and memory of the initial conditions will be evident in the


single-site magnetization. When starting in the Néel ordered state, a 0.75
Fig. 2b–f shows the time evolution of hσiz i for different disorder
strengths. The frozen moments of the spins increase with increasing
disorder as the emergent integrals of motion become more strongly 0.50
localized10 .
HD, (t)

With the maximum applied disorder, W = 8Jmax , we measure


the single-spin reduced density matrix for the initial state and
the averaged matrix for Jmax t ≥ 5. In this case, localization of the 0.25
spins leads to a marked difference in the measured and thermal W = 0Jmax W = 1Jmax
reduced density matrices, indicating memory of the system’s initial W = 2Jmax W = 4Jmax
W = 6Jmax W = 8Jmax
conditions and a breakdown of ETH.
To quantify the localization, we measure the normalized 0.00
0 2 4 6 8 10
Hamming distance (HD)22 : Jmaxt

1 1 X b 0.50
c 0.35
D(t) = − hψ0 | σiz (t)σiz (0) |ψ0 i (2) W = 8Jmax
Steady-state 〈 (t)〉

Steady-state 〈 (t)〉

2 2N i
0.33
0.31
which gives the number of spin flips away from the initial state, 0.40
ψ0 , normalized by the length of the chain, N . At long times, 0.29
the HD approaches 0.5 for a thermalizing state and remains at 0 0.27
for a fully localized state. In Fig. 3a, we measure that the long- 0.30 0.25
0.1 1 10 1.0 1.2 1.4 1.6 1.8
time HD is 0.5 in the absence of disorder, and becomes smaller
Disorder strength, W (Jmax) Power-law exponent, α
as the disorder strength is increased and the system more strongly
localizes. This time evolution agrees with exact diagonalization
numerical calculations for (1) (see Supplementary Information). Figure 3 | The Hamming distance (HD). a, The HD exhibits time dynamics
Figure 3b shows that for finite but weak disorder, the time- that reach their steady-state values after Jmax t ≈ 5. For increasing disorder,
averaged HD for Jmax t > 5 is essentially unchanged, indicating weak the system becomes more strongly localized, and the steady-state HD
or no localization. However, once the random field is sufficiently decreases. (Different colours represent different disorder strengths.) b, The
strong we observe a crossover from a thermalizing to a localized steady-state HD with respect to the strength of the random potential
state. Once in this regime, the system becomes more localized with indicates the state is not or only weakly localized for small disorder, but
increasing disorder strength. after the random field is sufficiently strong it becomes more localized with
There is great theoretical interest in mapping the MBL increased disorder. c, The steady-state HD with respect to the power-law
phase diagram with respect to interaction range and disorder exponent indicates that the system becomes less localized in the presence
strength11,22,23 . We have taken the first steps towards this goal of longer-range interactions (smaller α). Error bars, 1 s.d.

NATURE PHYSICS | VOL 12 | OCTOBER 2016 | www.nature.com/naturephysics 909


© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
LETTERS NATURE PHYSICS DOI: 10.1038/NPHYS3783

a b Data availability. The data that support the plots within this paper
1.5 No disorder W = 6Jmax
1.1 and other findings of this study are available from the corresponding
1.4 author upon reasonable request.

QFI
QFI

1.3 1.0 Received 11 December 2015; accepted 9 May 2016;


1.2
published online 6 June 2016
0.9
1.1 W = 8Jmax References
1. Kinoshita, T., Wenger, T. & Weiss, D. S. A quantum Newton’s cradle. Nature
0.3 1 3 10 0.3 1 3 10
440, 900–903 (2006).
Jmaxt Jmaxt
2. Gring, M. et al. Relaxation and prethermalization in an isolated quantum
system. Science 337, 1318–1322 (2012).
Figure 4 | Quantum Fisher information (QFI). a, Time evolution of the QFI 3. Anderson, P. W. Absence of diffusion in certain random lattices. Phys. Rev. 109,
for no disorder, which is consistent with no long-time growth of 1492–1505 (1958).
entanglement. The greyed-out area indicates the fast initial growth of QFI 4. Wiersma, D. S., Bartolini, P., Lagendijk, A. & Righini, R. Localization of light in
that follows a Lieb–Robinson-type bound. b, Long-time logarithmic growth a disordered medium. Nature 390, 671–673 (1997).
5. Billy, J. et al. Direct observation of Anderson localization of matter waves in a
of the QFI for the applied disorder of W = (6, 8)Jmax is a lower bound for the
controlled disorder. Nature 453, 891–894 (2008).
entanglement in the system and is consistent with the expected long-time 6. Roati, G. et al. Anderson localization of a non-interacting Bose–Einstein
growth of entanglement in the MBL state. Black lines are logarithmic fits to condensate. Nature 453, 895–898 (2008).
the data. Statistical error bars (1 s.d.) are smaller than the data points. 7. Basko, D. M., Aleiner, I. L. & Altshuler, B. L. Possible experimental
manifestations of the many-body localization. Phys. Rev. B 76, 052203 (2007).
8. Oganesyan, V. & Huse, D. A. Localization of interacting fermions at high
temperature. Phys. Rev. B 75, 155111 (2007).
by measuring a change in the time-averaged HD for W = 8Jmax 9. Pal, A. & Huse, D. A. Many-body localization phase transition. Phys. Rev. B 82,
and Jmax t > 5 as we adjusted the interaction range, 0.95 < α < 1.81 174411 (2010).
(Fig. 3c). For shorter-range interactions, the system appears more 10. Serbyn, M., Papić, Z. & Abanin, D. A. Local conservation laws and the
localized, because the state approaches a fully localized Anderson structure of the many-body localized states. Phys. Rev. Lett. 111, 127201 (2013).
insulator as α → ∞. This change in time-averaged HD with 11. Yao, N. et al. Many-body localization in dipolar systems. Phys. Rev. Lett. 113,
243002 (2014).
a change in interaction range makes clear that the long-range
12. Kondov, S., McGehee, W., Xu, W. & DeMarco, B. Disorder-induced localization
couplings are playing a role in the observed dynamics, thus in a strongly correlated atomic hubbard gas. Phys. Rev. Lett. 114, 083002 (2015).
indicating the observed effect is a many-body phenomenon. 13. Schreiber, M. et al. Observation of many-body localization of interacting
Although there are predictions of a many-body delocalization fermions in a quasi-random optical lattice. Science 349, 842–845 (2015).
transition at α = 1.5 (ref. 23), we did not observe this effect as we 14. Mølmer, K. & Sørensen, A. Multiparticle entanglement of hot trapped ions.
tuned α across this boundary. The lack of a sharp transition, along Phys. Rev. Lett. 82, 1835–1838 (1999).
15. Islam, R. et al. Emergence and frustration of magnetism with variable-range
with the presence of MBL states for α < 1, may be due to finite size interactions in a quantum simulator. Science 340, 583–587 (2013).
effects (see Supplementary Information). As this system is scaled 16. Wu, Y.-L. & Das Sarma, S. Understanding analog quantum simulation
to many dozens of spins, it will allow better study of the phase dynamics in coupled ion-trap qubits. Phys. Rev. A 93, 022332 (2016).
transition and mapping of the phase boundary in a regime where 17. Richerme, P. et al. Non-local propagation of correlations in quantum systems
numerics are intractable. with long-range interactions. Nature 511, 198–201 (2014).
18. Jurcevic, P. et al. Quasiparticle engineering and entanglement propagation in a
A hallmark of MBL is the characteristic growth of entanglement
quantum many-body system. Nature 511, 202–205 (2014).
under coherent time evolution24 , although its experimental 19. Deutsch, J. M. Quantum statistical mechanics in a closed system. Phys. Rev. A
observation has been elusive so far. In Anderson insulators without 43, 2046–2049 (1991).
many-body interactions, the entanglement production from weakly 20. Srednicki, M. Chaos and quantum thermalization. Phys. Rev. E 50,
entangled initial states shows a quick saturation after a sharp 888–901 (1994).
transient regime. However, in MBL systems a long-time growth sets 21. Rigol, M., Dunjko, V. & Olshanii, M. Thermalization and its mechanism for
generic isolated quantum systems. Nature 452, 854–858 (2008).
in, which is logarithmically slow for short-range interactions25 and 22. Hauke, P. & Heyl, M. Many-body localization and quantum ergodicity in
can become algebraic with power-law interactions26 . disordered long-range Ising models. Phys. Rev. B 92, 134204 (2015).
This entanglement growth can be measured using a suitable wit- 23. Burin, A. L. Localization in a random XY model with long-range interactions:
ness operator or even full state tomography27 . We instead indirectly intermediate case between single-particle and many-body problems.
characterize the entanglement growth in this system by measuring Phys. Rev. B 92, 104428 (2015).
the quantum Fisher information (QFI)28–30 . The QFI gives a lower 24. Nandkishore, R. & Huse, D. A. Many-body localization and thermalization
in quantum statistical mechanics. Annu. Rev. Condens. Matter Phys. 6,
bound on the entanglement in the system while requiring only 15–38 (2015).
a measurement of two-body correlators, which can be efficiently 25. Bardarson, J. H., Pollmann, F. & Moore, J. E. Unbounded growth of
accessed with our site-resolved imaging. Importantly, the QFI is able entanglement in models of many-body localization. Phys. Rev. Lett. 109,
to distinguish MBL from single particle localization through the 017202 (2012).
anticipated characteristic entanglement growth (see Supplementary 26. Pino, M. Entanglement growth in many-body localized systems with
long-range interactions. Phys. Rev. B 90, 174204 (2014).
Information). With no applied disorder, we observe a fast initial
27. Haeffner, H. et al. Scalable multiparticle entanglement of trapped ions. Nature
growth of the QFI following a Lieb–Robinson bound17,18 as the 438, 643–646 (2005).
correlations propagate through the system, but no further growth 28. Helstrom, C. W. Quantum Detection and Estimation Theory (Academic, 1976).
afterwards (Fig. 4a). In contrast, for the cases of applied disorder of 29. Braunstein, S. L. & Caves, C. M. Statistical distance and the geometry of
W = 6Jmax and W = 8Jmax , the further growth of the QFI is consistent quantum states. Phys. Rev. Lett. 72, 3439–3443 (1994).
with a logarithmic increase of entanglement at long times in a MBL 30. Pezzé, L. & Smerzi, A. in Proc. Int. School Phys. (eds Tino, G. & Kasevich, M.)
691–741 (IOS Press, 2014).
state (Fig. 4b), but absent for single particle localized systems.
Our experimental platform is well suited for studying deep
and intractable questions about thermalization and localization in Acknowledgements
quantum many-body systems. Moreover, the high degree of control We thank L. Duan, Z.-X. Gong, T. Grover, C. Laumann, S. Wang, N. Yao and J. Zhang for
helpful discussions. This work is supported by the ARO Atomic and Molecular Physics
in our experiment may guide the use of MBL states as potential Program, the AFOSR MURI on Quantum Measurement and Verification, the IARPA
quantum memories in naturally disordered quantum systems24 . MQCO Program, and the NSF Physics Frontier Center at JQI. M.H. acknowledges the

910 NATURE PHYSICS | VOL 12 | OCTOBER 2016 | www.nature.com/naturephysics

© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ


NATURE PHYSICS DOI: 10.1038/NPHYS3783 LETTERS
Deutsche Akademie der Naturforscher Leopoldina (grant No. LPDS 2013-07 and LPDR Additional information
2015-01) and P.H. acknowledges the EU IP SIQS, the SFB FoQuS (FWF Project Supplementary information is available in the online version of the paper. Reprints and
No. F4016-N23) and the ERC synergy grant UQUAM. permissions information is available online at www.nature.com/reprints.
Correspondence and requests for materials should be addressed to J.S.
Author contributions
J.S., A.L., P.R., B.N., P.W.H. and C.M. performed the experimental work and P.H., M.H.
and D.A.H. performed the theoretical work. All authors contributed to the preparation of Competing financial interests
this manuscript. The authors declare no competing financial interests.

NATURE PHYSICS | VOL 12 | OCTOBER 2016 | www.nature.com/naturephysics 911


© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
SUPPLEMENTARY INFORMATION
DOI: 10.1038/NPHYS3783

1 Many-body
Methods for:localization
Many-bodyin a quantuminsimulator
localization with
a quantum simu-
2 programmable random disorder
lator with programmable random disorder

3 J. Smith∗, A. Lee∗ , P. Richerme†, B. Neyenhuis∗ , P. W. Hess∗ , P. Hauke‡§, M. Heyl‡§¶, D. A. Huse,

4 C. Monroe∗

5 Generating the effective Hamiltonian We generate spin-spin interactions by applying spin-dependent

6 optical dipole forces to ions confined in a 3-layer linear Paul trap with a 4.8 MHz radial frequency.

7 Two off-resonant laser beams with a wavevector difference ∆k along a principal axis of transverse

8 motion globally address the ions and drive stimulated Raman transitions. The two beams contain a

9 pair of beatnote frequencies symmetrically detuned from the resonant transition at ν0 = 12.642819

10 GHz by a frequency µ, comparable to the transverse motional mode frequencies. In the Lamb-

11 Dicke regime, this results in the Ising-type Hamiltonian in Eq. (1) 1–3 with

N
2 bi,m bj,m
Ji,j = Ω ωR 2 , (2)
m=1
µ2 − ωm

12 where Ω is the global Rabi frequency, ωR = h̄∆k 2 /(2M ) is the recoil frequency, bi,m is the normal-

13 mode matrix 4 , and ωm are the transverse mode frequencies. The coupling profile may be approx-

14 imated as a power-law decay Ji,j ≈ Jmax /|i − j|α , where in principle α can be tuned between 0

15 and 3 by varying the laser detuning µ or the trap frequencies ωm 2, 5 . In this work, α was tuned ap-

Joint Quantum Institute, University of Maryland Department of Physics and National Institute of Standards and
Technology, College Park, MD 20742

Department of Physics, Indiana University, Bloomington, IN, 47405

Institute for Quantum Optics and Quantum Information of the Austrian Academy of Sciences, 6020 Innsbruck,
Austria
§
Institute for Theoretical Physics, University of Innsbruck, 6020 Innsbruck, Austria

Physik Department, Technische Universitt Mnchen, 85747 Garching, Germany

Physics Department, Princeton University, Princeton, NJ 08544, USA

1
NATURE PHYSICS | www.nature.com/naturephysics 1
© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

16 proximately between 0.95 and 1.81 by changing µ. By asymmetrically adjusting the laser beatnote

17 detuning µ about the carrier by a value of B we apply a global Stark shift that can be thought of as

18 a uniform effective transverse magnetic field of (B/2)σiz .

19 We generate the effective disorder by applying a site-dependent Stark shift using a single

20 355nm laser beam that is focused down tightly to a 1/e2 waist of ∼ 1.8µm. The ion separation is

21 ∼ 2.5µm, thus the crosstalk between ions is negligible with a measured ratio of nearest-neighbor

22 Rabi frequencies of ∼ 20 : 1. We use an acousto optic modulator (AOM) with a full width at half

23 maximum bandwidth of ≈100 MHz to apply the Stark shift to each ion. The AOM is not imaged

24 onto the ions, so that driving the AOM with different frequencies allows the position of the beam to

25 be scanned over the length of a 10 ion chain, ∼ 20µm. The Stark shift is proportional to I 2 . Thus,

26 to achieve larger applied Stark shifts, we raster through the AOM drive frequencies corresponding

27 to addressing each ion with a total cycle time of ∼ 5µs instead of applying them simultaneously.

28 Since we cannot control the sign of the site-specific Stark shift, to center the disorder strength

29 around the global transverse field, we adjust the asymmetric detuning by W Jmax /2.

30 Sampling Error We determine that averaging over 30 different random realizations of disorder is

31 sufficient to have a sampling error smaller than the effect we observe by looking at the change in

32 the time-averaged HD with respect to a change in the disorder strength. Figure 3b makes explicit

33 that this error is much smaller than than the change in the time-averaged HD with respect to a

34 change in the disorder strength.

2 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

35 Measuring the spin-spin coupling matrix In order to observe the dynamics between just two

36 of the ions in the chain, we shelve the other spins out of the interaction space. This is done

37 by performing a π rotation between |↓z , 2 S1/2 |F = 0, mF = 0, and one of the Zeeman states,
2
38 S1/2 |F = 1, mF = −1, while shifting the two ions of interest out of resonance by applying a

39 large Stark shift with the individual addressing beam. We then apply our Hamiltonian which now

40 acts only on the two ions left in the interaction space and determine the elements of the spin-spin

41 coupling matrix by fitting the measured interaction Rabi flopping frequency between each pair of

42 spins.

43 Arbitrary product state preparation State initialization starts with optically pumping the spins

44 with high-fidelity to |↓↓↓ · · · z . Then we perform a global π/2 rotation to bring the ions to

45 |↓↓↓ · · · x . At this point we apply a Stark shift with the individual addressing beam to the spins

46 that are to be flipped and allow the chain to evolve until these ions are π out of phase with rest

47 of the ions. This, along with our ability to perform high fidelity global rotations, allows for the

48 preparation of any arbitrary product state along any direction of the Bloch sphere. Individual spin

49 flips can be achieved with a fidelity of ∼ 0.97, while arbitrary state preparation can be done with a

50 fidelity of ∼ (0.97)N , where N is the number of spins flipped with the individual addressing beam.

51 Determining a Set of Thermalizing Parameters Extended Data Figure 1 shows the time evolu-

52 tion of σix  for different values of B for the spins initialized in the randomly chosen product state

53 |↓↓↓↑↓↓↓↑↓↑x . Without a transverse field, the spins are in an eigenstate of the Ising interaction

54 and undergo no evolution. Once a transverse field is added the individual spins begin to lose mem-

NATURE PHYSICS | www.nature.com/naturephysics 3


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

55 ory of their initial conditions and as its strength is increased, the ions thermalize faster and more

56 robustly.

57 To confirm the system is thermalizing, we measure the time evolution of the single site mag-

58 netization, σiz , along an orthogonal direction for different strengths of the transverse magnetic

59 field starting with the spins initialized in the Néel ordered state. As seen in Extended Data Fig. 1

60 the spins have lost information about their initial conditions in the z direction for all values of B.

61 We calculate the spectral statistics of adjacent energy levels for the Hamiltonian and find they

62 are not Poisson distributed for B = 4Jmax and Di = 0 indicating that with no applied disorder, the

63 system is not in a localized phase. Furthermore, one can determine if a system is in a thermal or

64 localized regime by finding the correlation between adjacent energy splittings by calculating the

65 ratio of two consecutive gaps6 :


min{δn , δn−1 }
rn = (3)
max{δn , δn−1 }

66 where δn = En+1 − En ≥ 0. For a localized phase, where one expects a Poisson energy spectrum,

67 the probability distribution of this order parameter is given by Pp (r) = 2/(1 + r)2 and thus r ≈

68 0.39. For energy level spacings following a random-matrix as predicted for a thermalizing regime,

69 we calculate r ≈ 0.53 for a chain of 10 spins. Extended Data Figure 2 shows that r saturates

70 to the expected value for a random matrix distribution, indicating that the Hamiltonian is thermal

71 for sufficiently large B.

4 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

72 Calculating the density matrix expected by the Eigenstate Thermalization Hypothesis Given

73 a Hamiltonian and an initial state |ψ0 , the corresponding energy is ψ0 | H |ψ0 . For a thermalizing

74 system satisfying ETH this energy should be equal to the classical energy:

Tr[He−βH ]
E= (4)
Tr[e−βH ]

75 for the appropriate β = 1/(kB T ). When partitioning the entire system into subsystems A and B,

76 with the size of A much smaller than B (perhaps even a single spin), then, the density matrix on

77 site A at long times can be approximated by:

TrB [e−βH ]
ρA = (5)
Tr[e−βH ]

78 Since we start in the Néel ordered state, the initial energy given the Hamiltonian in Eq. (1)

79 is equal to zero, ψ0 | H |ψ0  = 0. Equating this to the right hand side of Eq. (4) and solving for

80 β gives β = 0, or T = ∞. Using this β in Eq. (5) gives a value for any reduced thermal density

81 matrix of:  
 1/2 0 
 
 
0 1/2
82 in agreement with the measured reduced density matrices in Fig. 2a.

83 Comparison to Numerics To demonstrate the MBL we observe is a general feature of our Hamil-

84 tonian we perform numerical simulations using exact diagonalization. Extended Data Figure 3

85 compares the experimentally measured time evolution of the normalized HD with numerics and

86 shows excellent agreement between them. We see similar agreement between experimental data

87 and numerics for the time evolution of the single-spin magnetizations (not shown). The aspects of

NATURE PHYSICS | www.nature.com/naturephysics 5


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

88 MBL we experimentally measure were independently verified numerically as generic characteris-

89 tics of (1)7 .

90 Decoherence and Dephasing To measure our system’s coupling to the environment we fit an

91 exponential decay to the dynamics in the upper left panel of Extended Data Fig. 1 as we expect

92 no time evolution of σix  because the initial state is an eigenstate of the Hamiltonian and thus any

93 dynamics give an estimate of the decoherence rate. We find this estimate to be JM ax t = 64.6 which

94 is approximately 60 times slower than the dynamics of the localization.

95 Extended Data Fig. 4 shows a numerical simulation of the extended dynamics for the model

96 Hamiltonian with (green curves) and without (blue curves) crosstalk error between ions from the

97 individual addressing and laser intensity noise on the Ising couplings. It is clear that the localization

98 persists well beyond the experimental timescales even when accounting for experimental noise.

99 We model the crosstalk noise on the disordered field by adding 5% of the Stark shift applied

100 to adjacent ions to the size of the intended Stark shift which is consistent with the spillover we

101 measure between ions. To incorporate noise on the spin-spin couplings, we scale the strength of

102 the Ising couplings by a value we pull randomly from a Gaussian distribution centered around

103 µ = 1 with σ = 0.05 for each instance of disorder because the laser intensity noise is slower

104 than the duration of an experiment. The size of this simulated noise is consistent with the directly

105 measured noise on Ji,j .

106 Quantum Fisher Information The quantum Fisher information (QFI) has recently been shown
8, 9
107 to witness genuinely multipartite entanglement . From a quantum metrology perspective, the

6 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

108 QFI quantifies the sensitivity of a given input state to a unitary transformation eiϑÔ generated by

109 the hermitian operator Ô. In a pure state, it is given by 10

FQ = 4(∆Ô)2 = 4(Ô2  − Ô2 ). (6)

N
110 For a local operator Ô = i=1 Ôi (where the difference between largest and smallest eigen-

111 value of Ôi is 1), the QFI witnesses entanglement as soon as

fQ ≡ FQ /N > 1 . (7)

112 To characterize the growth of entanglement out of the initial Néel state, the natural choice
N i z
113 of the generator Ô is the staggered magnetization, Ô = i=1 (−1) σi /2. Remarkably, this QFI is

114 proportional to the variance of the HD D(t) given by Eq. (2) of the main text,

 
FQ = 4N 2 (∆D̂)2 = [(−1)i+j σiz σjz ] − [ (−1)i σiz ]2 , (8)
i,j i

N i z
115 when associating D(t) = D̂(t), with D̂ = 1/(2N )[1 − i=1 (−1) σi ].

116 The QFI as defined in Eq. (6) assumes a pure state, i.e., that time evolution is purely unitary.

117 For mixed states, the QFI cannot be expressed as a simple expectation value of the operator Ô10 .

118 In general, decoherence reduces the purity of the system’s state over experimental time scales. To

119 show that the measured increase of FQ as defined in Eq. (8) is indeed due to coherent dynamics, we

120 compare to numerical calculations for a unitary time evolution using the experimental parameters.

121 Extended Data Fig. 5 shows the experimental data is always below the theoretical prediction for

NATURE PHYSICS | www.nature.com/naturephysics 7


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

122 a unitary time evolution. The loss of purity or other experimental imperfections such as detection

123 error, therefore, do not generate a false positive indicator of entanglement in our system.

124 Extended Data Figure 6 further establishes this point, showing an increase in the QFI we

125 measure and strong agreement between experimental data and theory when postselecting for mea-

126 sured states with 5-spin excitations. We postselect because when B  J the Hamiltonian is

127 effectively an XY model and conserves i σiz , because Ising processes that flip spins along the

128 large field are energetically forbidden. However, because of camera detection error we find there is

129 significant leakage out of states with 5-spin excitations (≈ 70% expected numerically, ≈ 35% de-

130 tected) into states with 4 and 6-spin excitations (≈ 20% detected) which should not be populated

131 as the transverse magnetization is conserved modulo two spin flips in the transverse field Ising

132 model. Thus, we post-select for states with 5-spin excitations. Extended Data Figures 5 and 6

133 show a clear difference between the interacting case and a theory of free-fermions (see below) for

134 the experimental data and numerical simulations, thus, establishing that the growth in QFI in the

135 data and full-Hamiltonian numerics are due to a many-body effect.

136 To study how the localization changes with system size, we performed a numerical finite-size

137 scaling. In order to obtain a well-behaved scaling, we use the Kac prescription11 , i.e., we adjust

138 the couplings as Jij = JN −1 |i − j|−α , where N = (N − 1)−1 i<j |i − j|−α . Note that using

139 this prescription the fundamental energy scale J differs by about a factor of 2 from Jmax , the value

140 used in the main text.

141 For α > 2, the disordered long-range Ising Hamiltonian shows MBL behavior at large

8 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

142 disorder12 . In Extended Data Fig. 7, we plot the dynamics of the QFI for α = 3, where it grows

143 consistent with the characteristic long-time growth of entanglement for an MBL state. In par-

144 ticular, within a time window 2α < tJ < 3α where only next-nearest-neighbor interactions are

145 relevant, the system essentially behaves as a nearest-neighbor Ising model with a weak next-to-

146 nearest-neighbor coupling. For such a system, a logarithmic growth of entanglement is expected,

147 as we indeed find in that regime, see inset in Extended Data Fig. 7.

148 Moreover, in Extended Data Fig. 7, we compare our numerical results to the appropriate

149 long-range free-fermionic theory (see below), which shows a quick system-size independent sat-

150 uration of the QFI without further growth. Therefore, we conclude that the observed increase of

151 the QFI is not possible in a quantum system without many-body interactions, thus giving a clear

152 signature for true MBL behavior.

153 The situation is more complex at α = 1.13. For B = 0, it has been predicted that within

154 the range 1 < α < 2 delocalized behavior could be expected in the thermodynamic limit12 . As

155 seen in Extended Data Fig. 8, for the considered system sizes up to N = 14 the model displays all

156 essential signatures of MBL, as found for α = 3. However, the important question of whether this

157 localization persists in the thermodynamic limit can only be addressed with system sizes larger

158 than those accessible using exact diagonalization. Here, scaling our quantum simulator to larger

159 system sizes could thus resolve a difficult open question, namely of the existence of ergodicity in

160 the range 1 < α < 2. However, we would like to emphasize that the essential features of MBL

161 are nevertheless captured by the 10-spin experimental system. In particular, we still find a time

NATURE PHYSICS | www.nature.com/naturephysics 9


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

162 window consistent with a logarithmic growth of entanglement, see inset in Extended Data Fig. 8.

163 In order to make a stronger connection between growth in the QFI and growth of entangle-

164 ment we calculate the entropy of entanglement between two halves of the chain:

SA = −T r[ρA log ρA ] (9)

165 where ρA = T rB [ρ] and B is the other half of the spin chain. The entanglement entropy quantifies

166 the number of entangled bits between two subsystems.

167 In Extended Data Figs. 7 and 8 it is clear that there is long-time growth of the entanglement

168 entropy that is consistent with the expected growth for a MBL state13, 14 and is absent in the free-

169 fermion numerics. The difference between the numerics for the model Hamiltonian and the non-

170 interacting theory for the QFI and the entanglement entropy in Extended Data Figs. 9 and 10

171 distinguishes between the two cases for the experimental system size and timescale. These figures

172 also establish a qualitative connection between growth in the QFI and growth in entanglement.

173 To show that the QFI growth is truly due to interactions, we also compare numerics with the

174 experimentally applied Hamiltonian to a close approximation of H, Eq. (1) of the main text, with a

175 non-interacting theory. Using the Jordan-Wigner transformation, σj− → e−iθj cj , with the phase of
 †
176 the string operator θj = π j<i cj cj , the Hamiltonian Eq. (1) can be mapped to a fermionic theory

177 with annihilation and creation operators cj and c†j , respectively,

 
H= Jij (c†i ei(θj −θi ) cj + c†i ei(θj +θi ) c†j + h.c.) − (B + Di )c†i ci . (10)
i<j i

178 If Jij contained only nearest-neighbor interactions, this Hamiltonian would become equivalent

10

10 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

179 to a free-fermionic theory. For general Jij , however, the string operators generate interactions

180 between the fermions. Over short times, and especially in a localized regime, the phases θj are

181 dominated by their initial values, i.e., it is a good approximation to replace (for the initial Neel
 j
182 state) θj → π j<i ((−1) + 1)/2 in the Hamiltonian. This replacement amounts to approximating

183 H by a non-interacting fermionic theory with long-range hopping and pairing. The QFI for that

184 case is included in Extended Data Figs. 7 and 8. As one can see, the QFI quickly saturates to

185 values below fQ = 1. The experimentally and numerically observed further growth of the QFI

186 is thus truly due to interactions, and cannot be captured within a free-fermionic theory, even with

187 long-range hopping.

189
188 1. Mølmer, K. & Sørensen, A. Multiparticle Entanglement of Hot Trapped Ions. Phys. Rev. Lett.

190 82, 1835-1838 (1999).

191 2. Porras, D. & Cirac, J. I. Effective Quantum Spin Systems with Trapped Ions. Phys. Rev. Lett.

192 92, 207901 (2004).

193 3. Kim, K. et al. Entanglement and Tunable Spin-Spin Couplings between Trapped Ions Using

194 Multiple Transverse Modes. Phys. Rev. Lett. 103, 120502 (2009).

195 4. James, D. F. V. Quantum dynamics of cold trapped ions with application to quantum compu-

196 tation. Applied Physics B 66, 181-190 (1998).

197 5. Islam, R. et al. Emergence and Frustration of Magnetism and Variable-Range Interactions in

198 a Quantum Simulator. Science 340, 583-587 (2013).

11

NATURE PHYSICS | www.nature.com/naturephysics 11


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

199 6. Oganesyan, V. & Huse, D. A. Localization of interacting fermions at high temperature. Phys.

200 Rev. B 75, 155111 (2007).

201 7. Wu, Y-L. & Das Sarma, S. Understanding analog quantum simulation dynamics in coupled

202 ion-trap qubits. Phys. Rev. A 93, 022332 (2016).

203 8. Hyllus, P. et al. Fisher information and multiparticle entanglement. Phys. Rev. A 85, 022321

204 (2012).

205 9. Tóth, G. Multipartite entanglement and high precision metrology. Phys. Rev. A 85, 022322

206 (2012).

207 10. Braunstein, S. L. & Caves, C. M. Statistical distance and the geometry of quantum states.

208 Phys. Rev. Lett. 72, 3439-3443 (1994).

209 11. Kac, M. Can One Hear the Shape of a Drum? Am. Math. Monthly 73, 1-23 (1966).

210 12. Burin, A. L. Energy delocalization in strongly disordered systems induced by the long-range

211 many-body interaction. arXiv:cond-mat/0611387 (2006).

212 13. Bardarson, J. H., Pollmann, F. & Moore, J. E. Unbounded Growth of Entanglement in Models

213 of Many-Body Localization. Phys. Rev. Lett. 109, 017202 (2012).

214 14. Pino, M. Entanglement growth in many-body localized systems with long-range interactions.

215 Phys. Rev. B 90, 174204 (2014).

12

12 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

B=0Jmax B=1Jmax B=4Jmax B=8Jmax


1.0
X Magnetization, 〈σix〉

0.5

0.0

-0.5

-1.0

1.0
Z Magnetization, 〈σiz〉

0.5

0.0

-0.5

-1.0
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
216
Jmaxt Jmaxt Jmaxt Jmaxt

217 Extended Data Figure 1: Measured thermalization in the transverse field Ising model. The

218 upper panels show the time dynamics of σix  (different colors represent different ions) for 10

219 spins prepared in the random product state |↓↓↓↑↓↓↓↑↓↑x , for different transverse magnetic field

220 strengths. For B = 0 the spins are in a eigenstate and do not thermalize. However, as the strength

221 of B is increased the system begins to thermalize more robustly and quickly. The lower panel plots

222 the time evolution of σiz  with 10 spins prepared in the Néel ordered in the z direction for different

223 transverse magnetic field strengths. We conclude that the system is in the thermalizing regime for

224 B = 4Jmax since we observe thermalizing behavior along two orthogonal directions. Error bars

225 are 1 standard deviation of statistical error.

13

NATURE PHYSICS | www.nature.com/naturephysics 13


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

0.54
  

0.52 


0.50
r 



0.48 

0.46
1 2 5 10 20 50
226 BJmax

227 Extended Data Figure 2: Thermalizing level statistics. The calculated value of r with respect

228 to B saturates close to the predicted value for a random-matrix distribution (dashed black line)

229 implying that the Hamiltonian is in the thermal phase for sufficiently large B.

14

14 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

0.6

0.4
D(t)

0.2
W/J max =8
W/J max =6
W/J max =4
W/J max =2
0
0 2 4 6 8 10
230 Jmax t

231 Extended Data Figure 3: Comparison of the experimental data (crosses) with exact numerical

232 simulations (blue lines) for normalized Hamming distance. There is excellent agreement be-

233 tween the numerical simulations using the experimental parameters and the measured data. This

234 demonstrates that the observed effects are general features of the Hamiltonian.

15

NATURE PHYSICS | www.nature.com/naturephysics 15


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

0.6

0.4
D(t)

0.2 W/J max =8


W/J max =6
W/J max =4
W/J max =2
0
0.1 1 10 100 500
235
Jmax t

236 Extended Data Figure 4: Numerical simulations of the extended time evolution of the Ham-

237 ming distance for the model Hamiltonian (blue curves) and with experimental noise (green

238 curves). The localization we observe persists much longer than the experimental timescale in

239 the model Hamiltonian (blue curves) even when accounting for laser intensity noise and crosstalk

240 between the ions from the individual addressing beam (green curves).

16

16 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

1.4
W/J max = 6 W/J max =8
1.2
fQ (t)

0.8
0.1 1 10 0.1 1 10
241 Jmax t Jmax t

242 Extended Data Figure 5: Comparison of experimental data (green dots) with exact numerical

243 simulations for the experimentally applied Hamiltonian (thick blue lines) and free-fermion

244 Theory (orange) for QFI. The solid straight lines represent logarithmic fits to the numerical (light

245 blue) and experimental data (light green). Deviations from the ideal coherent dynamics due to

246 decoherence and other imperfections in the experimental setup, such as detection error, lead to a

247 reduction of the QFI. Importantly, this suggests that experimental imperfections do not generate a

248 false positive for entanglement. Moreover, there is long-time growth in the QFI from the measured

249 data and applied Hamiltonian numerics that is absent in the free-fermion theory.

17

NATURE PHYSICS | www.nature.com/naturephysics 17


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

1.4
W/J max = 6 W/J max = 8
1.2
fQ (t)

0.8
0.1 1 10 0.1 1 10
250 Jmax t Jmax t

251 Extended Data Figure 6: Comparison of postselected experimental data (green dots) with ex-

252 act numerical simulations for the experimentally applied Hamiltonian (thick blue lines) and

253 free-fermion Theory (orange) for QFI. The solid straight lines represent logarithmic fits to the

254 numerical results for the experimentally applied Hamiltonian (light blue) and postselected data

255 for results with 5 spin flips (light green). The increase in the postselected QFI and the agreement

256 between the postselected data and numerical simulations supports the claim that experimental im-

257 perfections decrease the value of the QFI for the full experimental data.

18

18 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

1.5 2

1
fQ (t)

S(t)
1
1.3

0.5
fQ (t)

1.1

0.9
1 10 100
0 0
0.1 1 10 100 1000 0.1 1 10 100 1000
258 Jt Jt

259 Extended Data Figure 7: QFI and entanglement entropy from exact diagonalization (α = 3

260 and W/J = 8). Left panel: When subject to disorder, the QFI of the staggered magnetization

261 shows a characteristic growth of entanglement (blue lines; from dark to light: N = 8, 10, 12, 14

262 averaged over 106 , 3 × 105 , . . . , 2500 disorder realizations). This growth is absent in a theory of

263 free-fermions with long-range hopping and pairing (green dashed lines with N = 14 (dark green)

264 averaged over 10000 realizations). Left panel inset: In a time window dominated by next-nearest

265 neighbor interactions, 2α < tJ < 3α , one observes a characteristic logarithmic entanglement

266 growth, expected for a MBL system with short-range interactions. Right panel: The entanglement

267 entropy between two halves of the chain shows long-time logarithmic growth for the interacting

268 case and saturates for the free-fermion theory consistent with the expectation for a MBL state and

269 single-particle localized state, respectively, and a qualitative agreement between growth in QFI and

270 entanglement entropy.

19

NATURE PHYSICS | www.nature.com/naturephysics 19


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

1.5 3

1 2
fQ (t)

S(t)
1.3

0.5 1
fQ (t)

1.1

0.9
1 10 100
0 0
0.1 1 10 100 1000 0.1 1 10 100 1000
271 Jt Jt

272 Extended Data Figure 8: QFI and entanglement entropy from exact diagonalization (α =

273 1.13 and W/J = 8). Same color coding as in Extended Data Fig. 7. Importantly, for the experi-

274 mentally relevant system size of N = 10, we again find a time window consistent with a logarithmic

275 growth of entanglement in the growth of QFI (see left inset) and half-chain entanglement entropy.

20

20 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.
DOI: 10.1038/NPHYS3783
SUPPLEMENTARY INFORMATION

0.25 1
L =8 L =8
L =10 L =10
0.2 L =12 L =12 0.8
L =14 L =14

0.15 0.6
δ fQ (t)

δ S(t)
0.1 0.4

0.05 0.2

0 0
0.1 1 10 100 1000 0.1 1 10 100 1000
276 Jt Jt

277 Extended Data Figure 9: Difference between numerics for the interacting model Hamiltonian

278 and free-fermion theory for QFI and entanglement entropy (α = 3 and W/J = 8). There is

279 a clear departure between the numerically calculated QFI and entanglement entropy for the model

280 Hamiltonian and the free-fermion theory.

21

NATURE PHYSICS | www.nature.com/naturephysics 21


© 2016 Macmillan Publishers Limited. All rights reserved.
SUPPLEMENTARY INFORMATION DOI: 10.1038/NPHYS3783

0.5 2
L =8 L =8
L =10 L = 10
L =12 L = 12
L =14 L = 14
δ fQ (t)

δ S(t)
0.25 1

0 0
0.1 1 10 100 1000 0.1 1 10 100 1000
281 Jt Jt

282 Extended Data Figure 10: Difference between numerics for the interacting model Hamiltonian

283 and free-fermion theory for QFI and entanglement entropy (α = 1.13 and W/J = 8).

284 There is a clear departure between the numerically calculated QFI and entanglement entropy for

285 the model Hamiltonian and the free-fermion theory for N=10 on the experimental timescale.

22

22 NATURE PHYSICS | www.nature.com/naturephysics


© 2016 Macmillan Publishers Limited. All rights reserved.

You might also like