Skip to main content
NIST Author Manuscripts logoLink to NIST Author Manuscripts
. Author manuscript; available in PMC: 2020 Dec 22.
Published in final edited form as: Phys Rev A (Coll Park). 2020;102:10.1103/PhysRevA.102.010401. doi: 10.1103/PhysRevA.102.010401

Signaling and scrambling with strongly long-range interactions

Andrew Y Guo 1,2, Minh C Tran 1,2,3, Andrew M Childs 1,4,5, Alexey V Gorshkov 1,2, Zhe-Xuan Gong 6,*
PMCID: PMC7754795  NIHMSID: NIHMS1642289  PMID: 33367192

Abstract

Strongly long-range interacting quantum systems—those with interactions decaying as a power law 1/rα in the distance r on a D-dimensional lattice for αD—have received significant interest in recent years. They are present in leading experimental platforms for quantum computation and simulation, as well as in theoretical models of quantum-information scrambling and fast entanglement creation. Since no notion of locality is expected in such systems, a general understanding of their dynamics is lacking. In a step towards rectifying this problem, we prove two Lieb-Robinson-type bounds that constrain the time for signaling and scrambling in strongly long-range interacting systems, for which no tight bounds were previously known. Our first bound applies to systems mappable to free-particle Hamiltonians with long-range hopping, and is saturable for αD/2. Our second bound pertains to generic long-range interacting spin Hamiltonians and gives a tight lower bound for the signaling time to extensive subsets of the system for all α< D. This many-site signaling time lower bounds the scrambling time in strongly long-range interacting systems.

Introduction.

In nonrelativistic quantum mechanics, Lieb-Robinson bounds provide a notion of causality [1], limiting the speed of information propagation (or signaling) to a finite value in lattice systems with short-range interactions. This bounded signaling speed has strong implications for quantum information and condensed matter physics, leading to entanglement area laws [2] and the existence of topological order [3]. However, it remains an open question whether the signaling speed must be finite if interactions are long-ranged and decay as an inverse power law 1/rα in the interparticle distance r. Such power-law interacting systems arise in experimental platforms for quantum computation and quantum simulation, including Rydberg atoms [4], trapped ions [5], polar molecules [6], defect centers in solids [7], and atoms trapped along photonic crystals [8]. The lack of a bounded signaling speed in these systems makes it challenging to understand and predict their dynamics.

For power-law interacting systems with α greater than the lattice dimension D, a finite signaling speed has been shown to persist to some intermediate distance and time [9]. At long distances (or times), recent developments show that the signaling speed will diverge at most polynomially in time for α > 2D [10,11], ruling out the exponential divergence suggested by earlier results [12]. For α ⩽ 2D, which is the case for most experimental long-range interacting systems, an exponentially growing signaling speed has yet to be ruled out, making the fate of causality far from settled.

In this work, we focus on the regime of strongly long-range interacting systems, where interaction energy per site diverges, thus implying αD [1316]. Note that even if one normalizes the interaction strength to make energy extensive (i.e., proportional to the number of lattice sites), these systems are still fundamentally different from those with α > D (as energy is in general no longer additive for subsystems [17]). To avoid confusion, we will not perform any normalization of interaction strength throughout this Rapid Communication, as such normalization can be performed later by rescaling time without changing the physics implied by our results [15].

Apart from their existence in experimental platforms [5,6,1820], strongly long-range interacting systems have received much theoretical interest due to their applications in spin squeezing [21], novel behavior in dynamical critical scaling [22,23], divergent equilibration time [13], and close relation to fast quantum-information scrambling [2430]. The phenomenology of these systems differs from that of their short-range counterparts at a fundamental level, and thus require new theoretical understandings. Two fundamental questions about these systems are (1) what is the shortest time tsi needed to send a signal from one site to a site located an extensive distance away, and (2) what is the shortest time tsc needed to scramble the information stored in the system [31]?

There have been a number of attempts to answer the above two questions, with limited success. For the first question, Refs. [3234] show that in certain strongly long-range interacting systems with αD, information and correlations can spread across the entire system in a finite time that is independent of the number of sites, N. (For certain systems not engineered for fast signaling or scrambling, information propagation may even be suppressed [35].) The Lieb-Robinsontype bound derived in Ref. [15], however, suggests that the signaling time can vanish in the N → ∞ limit, and does not rule out the possibility of tsi scaling as log(N)N2α/D/N2 for α < D. No protocol that we know of comes close to achieving such fast signaling. As for scrambling, Ref. [36] shows that the scrambling time can be lower-bounded by tsc ≿ 1/N for α = 0, whereas the fastest-known scramblers are conjectured to be able to scramble in time tsclog(N)/N [28].

While the definitive answers to these two questions remain to be found, we present several advances in this Rapid Communication. First, we prove a new bound for systems that can be mapped to free bosons or fermions with 1/rα hopping strength, which leads to a signaling-time bound of tsiNα/D/N. While no previous bound has been given specifically for free-particle systems, the best existing result for interacting systems yields a significantly looser bound of tsi ≿ log(N)N2α/D/N2 [15,37]. Notably, our free-particle bound is tight for αD/2, as we show that it can be saturated by a new quantum state transfer protocol.

We also prove a bound of tsi ≿ log(N)Nα/D/N for general interacting spin systems, which—while improving significantly over the previous best bound mentioned above [15]—is still not known to be tight. Building on this second result, we prove a tight bound for “many-site signaling” (from one site to an extensive part of the system). This many-site signaling bound leads to a scrambling-time bound of tscNα/D/N, which generalizes the result in Ref. [36] of tsc ≿ 1/N to all α < D.

Tight bound for free particles.

We first prove a Lieb-Robinson-type bound for noninteracting bosons/fermions on a lattice. Consider the following free-particle Hamiltonian H(t) defined on a D-dimensional lattice Λ with N sites:

H(t)=i,jΛi<j[Jij(t)cicj+H.c.]+iΛBi(t)cici, (1)

where ci (ci) represents the creation (annihilation) operator. The hopping strength Jij(t) and chemical potential Bi(t) can depend on time and we do not impose any constraint on them for now. We denote an operator A at time t in the Heisenberg picture as A(t) = U(t)AU(t), where U(t)Tei/0tH(t)dt is the time evolution operator (ℏ = 1). The operator norm of A will be denoted by ∥A∥.

Theorem 1.

For the Hamiltonian defined in Eq. (1) and any pair of distinct sites X, Y ∈ Λ,

[cX(t),cY]0tdτiΛ|JiX(τ)|2. (2)

We use [·,·] to denote the commutator for bosons and the anticommutator for fermions.

Roughly speaking, the quantity [cX(t),cY] measures the overlap between the support of the operator cX (t) (which expands from site X due to hopping terms) and the site Y. As a result, it also quantifies the amount of information that can be sent between X and Y in a given time t. Indeed, we define the signaling time tsi as the minimal time required to achieve [cX(t),cY]>δ for some positive constant δ. Note that we do not expect the chemical potential strength Bi(t) to show up in the bound, as on-site Hamiltonian terms do not change the support of cX (t).

If the hopping terms in the Hamiltonian are short-ranged (e.g., nearest-neighbor), one might expect [cX(t),cY] to decay exponentially in the distance rXY between X and Y, due to the strong notion of causality that follows from the Lieb-Robinson bound [1]. Additionally, if the hopping strength decays as a power law (|Jij(t)|1/rα) with α > D, intuition would suggest that [cX(t),cY] decays algebraically in rXY [9,12], indicating a weak notion of causality. However, the right-hand side of Eq. (2) has no dependence on rXY. This is because the bound is tailored to strongly long-range hoppings with α < D, which makes it loose for shorter-ranged long-range hoppings.

Assuming that |Jij(t)|1/rα, we can simplify Eq. (2) to

[cX(t),cY(0)]t×{O(1),α>D/2O(N1/2α/D),0αD/2, (3)

where N is the number of lattice sites and O is the asymptotic “big-O” notation [38]. Therefore, for αD/2, it takes a time tsi=Ω(Nα/D/N) [38] to signal from site X to site Y, independent of the distance between X and Y.

In the next section, we show that for αD/2, the bound in Eq. (2) can be saturated (up to a factor of 2) by engineered free-particle Hamiltonians. This leads to the conclusion that causality can completely vanish—in the sense that signals can be sent arbitrarily fast given large enough N—for a strongly long-range hopping system with α < D/2. It remains an open question whether such a statement can be generalized to systems with D/2 ⩽ α < D for either free or interacting particles.

Proof of Theorem 1.

Let us first go into the interaction picture of iBi(t)cici to eliminate the on-site terms from the Hamiltonian in Eq. (1). [This imparts a time-dependent phase eiϕjk(t) onto the hopping term Jjk(t) for some ϕjk(t) ∈ [0, 2π) and jk, which, since it does not change the value of |Jjk (t)|, does not affect the overall bound.] We now have a pure hopping Hamiltonian HI(t)=ijJ˜ij(t)cicj with |J˜ij(t)||Jij(t)|. Because HI (t) is a quadratic Hamiltonian, cX (t) is a time-dependent linear combination of annihilation operators on every site, and we can write [cX(t),cY]fXY(t)1, where fXY (t) is a number and 1 represents the identity operator. Given that UI(t)|0=|0, where UI (t) is the time-evolution operator corresponding to HI (t), and cX(t)|0=0, we have

fXY(t)=0|[cX(t),cY]|0=0|cX(0)UI(t)cY|0. (4)

For convenience, we define the (normalized) states |ψXcX(0)|0 and |ψY(t)UI(t)cY|0. Taking the time derivative of Eq. (4) gives

dfXYdt=iψX|HI(t)|ψY(t). (5)

By the Cauchy-Schwarz inequality,

dfXYdtHI(t)|ψX|ψY(t) (6)
=HI(t)|ψX=iΛ|J˜iX(t)|2. (7)

The last equality follows from |ψX ⟩ being a single excitation localized on site X and HI (t) consisting only of hopping terms J˜ij(t)cicj. Applying the fundamental theorem of calculus yields the bound on fXY (t) and hence Theroem 1.

Saturating the free-particle bound.

We now show that the bound in Theroem 1 can be saturated by engineered Hamiltonians that can also be used to perform fast quantum-state transfer. In particular, the protocol presented here has a state transfer time of T=O(Nα/D/N), which, for αD/2, improves over the fastest-known state transfer protocol using long-range interactions [34].

Our setup for the state transfer task is depicted in Fig. 1. We initialize a lattice with N sites in a tensor product of unoccupied states |0⟩ and some unknown normalized bosonic/fermionic state |ψ⟩ = a|0⟩ + b|1⟩ on a single site X. The goal of state transfer is to move |ψ⟩ to the target site Y after the system time-evolves by a |ψ⟩-independent (but possibly time-dependent) Hamiltonian H (t) [39,40].

FIG. 1.

FIG. 1.

A fast quantum-state transfer protocol for a long-range Hamiltonian acting on a lattice of dimension D = 1 with N = 7 sites. The strengths of the hopping terms are bounded by a power law 1/rα in the distance r. The active interactions in each time step are depicted as directed edges with uniform weights. (a) The site X is initially in the state |ψ⟩ (gray circle), with the other (unoccupied) sites in state |0⟩. Time-evolving by the Hamiltonian H1 for time O(Nα/D−1/2) (indicated by gray arrows) yields a superposition of the |0⟩N state and a symmetric |W⟩ state over the remaining N − 2 sites. (b) Applying the Hamiltonian H2 for the same duration of time completes the state transfer of |ψ⟩ to the target site Y.

The unitary time-evolution operator U (T) can be said to implement state transfer in time T if it satisfies the following condition:

|0|X0|N2ψ|YU(T)ψX|0N2|0Y|=1. (8)

We refer to the left-hand side of Eq. (8) as the fidelity of the state transfer, which can be bounded directly by a Lieb-Robinson-type bound on H (t) such as Eq. (2) [40].

We label the sites that are not X or Y by 1, …, N − 2 and denote the furthest distance between any pair of sites by L = O(N1/D). Our state transfer protocol is given by the following piecewise time-independent Hamiltonian:

H(t)={H1=1Lαi=1N2cXci+H.c.,0<t<T2,H2=1Lαi=1N2cicY+H.c.,T2<t<T, (9)

where T=πLα/N2 is the total time for the protocol. Note that while H (t) satisfies the constraint |Jij(t)|1/rijα assumed in Eq. (3), the corresponding Jij(t) terms do not actually vary with the distances between sites.

Evolving the initial state |Ψ|ψX|0N2|0Y by H1 for time T/2 yields the intermediate state

eiH1T/2|Ψ=a|0N+b|0X|W|0Y. (10)

Here, |W=1N2i=1N2ci|0N2 is the W state over the N − 2 remaining sites. Further evolving the state by H2 for time T/2 yields the final state:

eiH2T/2eiH1T/2|Ψ=|0X|0N2(a|0Y+b|1Y). (11)

Thus we have achieved perfect quantum-state transfer in time T=O(Nα/D/N). Note that the distance between X and Y on the lattice does not appear in the state transfer time. Setting b = 1 in the above protocol leads to

Ψ|[cX(T),cY]|Ψ=120TdτiΛ|JiX(τ)|2. (12)

Thus, the bound in Eq. (2) is saturated up to a factor of 2.

It should be pointed out that, for α > D/2, the above protocol requires a time that increases with N, which is slower than for the previous result in Ref. [34]. While that protocol has a state transfer time that is constant for αD, it uses an engineered Hamiltonian with interactions, and therefore cannot be applied to systems of noninteracting particles. In general, allowing interactions may increase the rate of information propagation, and proving a Lieb-Robinson-type bound in these situations requires a different approach.

Improved bound for general interacting systems.

We now derive bounds on the signaling time that extend beyond free-particle Hamiltonians. Without loss of generality, we study a generic interacting spin Hamiltonian H(t)=i<jhij(t) where hij(t)1/rijα and on-site interactions have been eliminated by going into an interaction picture. We will bound the quantity [A(t),B], where A and B are arbitrary operators supported on sets of sites X and Y, respectively, using the following Lieb-Robinson series [12]:

[A(t),B]2AB|XY|k=1(2t)kk!Jk(X,Y), (13)
Jk(X,Y)i1,,ik1JXi1Ji1i2Jik1Y. (14)

Here, |X| stands for the number of sites X acts on. Each term in Eq. (14) represents a sequence of k directed hops in the lattice that originates at site X and ends at site Y. For distinct sites i and j, Jij=1/rijα represents a directed hop from i to j. For technical reasons, we set Jii=jiJij [41].

Since Jij decays slowly in rij for α < D, our improved bound on [A(t),B] requires bounding each term in Eq. (14) using a new summation technique [42] absent in previous efforts [12,15]. This technique is particularly effective for tightening existing Lieb-Robinson bounds for strongly long-range interacting systems. The result (assuming α < D) is

[A(t),B]2AB|XY|(eΘ(N1α/D)t1Θ(N1α/D)rXYα). (15)

The factor Θ(N1α/D) [38] comes from the total interaction energy per site given by Jii.

We consider now the case of signaling between subsystems X and Y of a system Λ with |X|, |Y| = O(1). We formally define tsi, the signaling time from X to Y, as the smallest time t such that for a fixed constant δ = Θ(1), there exist unit-norm operators A and B supported on X and Y, respectively, such that [A(t),B]>δ [36]. If we further assume that X and Y are separated by an extensive distance rXY=Θ(N1/D), then the following lower bound holds for the signaling time:

tsi=Ω(log(N)N1α/D). (16)

This bound supersedes the naive signaling-time bound of tsi = Ω(1/N1−α/D) one would obtain via normalization of interaction energy per site. While we do not know of any examples that saturate this bound, it is the tightest-known signaling-time bound for strongly long-range interacting systems. Indeed, the bound is close to being saturated in the limit of αD, as the state transfer protocol in Ref. [34] shows that tsi = O(log N) can be achieved at α = D. Unfortunately, generalizing our bound in Eq. (15) to the case of α = D leads to tsi = Ω(1) [42], which is not saturated by Ref. [34].

Many-site signaling and scrambling bounds.

Of recent interest in the fields of theoretical high-energy and condensed matter physics has been the phenomenon of quantum information scrambling [28,36,4351]. Previous work on scrambling in power-law interacting systems has focused primarily on numerical analysis [52,53], whereas general mathematical results are lacking. Only in all-to-all interacting systems (which can be treated as the limit α = 0) have Lieb-Robinson-type bounds been used to bound the scrambling time [36]. Using the bound derived in Eq. (15), we can prove a scrambling-time bound for systems with 0 < α < D, a regime for which no better result is known.

To derive a bound on the scrambling time, we first derive a bound on the many-site signaling time. We define the “many-site signaling time” tms to be the smallest t required to signal from X to a Y that has extensive size. Lieb-Robinson-type bounds such as Eq. (15) naturally limit the time for many-site signaling. However, a direct application of Eq. (15) to many-site signaling leads to a loose bound. Instead, a more refined technique that sums over all sites within the subsets X and Yss yields a tighter bound [54]:

[A(t),B]2ABiX,jYeΘ(N1α/D)t1Θ(N1α/D)rijα. (17)

This bound reduces to Eq. (15) when |X|, Y| = 1.

The scrambling time tsc corresponds to the minimal time required for a system of N spins on a lattice Λ to evolve from a product state to a state that is nearly maximally entangled on all subsystems of size kN for some constant 0<k<12 [36]. From this definition, it can be shown that any information initially contained in a finite-sized subsystem S ⊂ Λ is no longer recoverable from measurements on S alone [36]. That information is not lost, however, but can be recovered from the complement S¯=Λ\S of S [44,55]. As a result, scrambling implies the ability to signal from S to S¯ [36]. Thus, tsc is lower bounded by the time it takes to signal from a subset S with size Θ(1) to its complement with size Θ(N), which corresponds to the many-site signaling time.

Using Eq. (17), we obtain the following scrambling-time bound for 0 ⩽ α < D:

tsctms=Ω(1N1α/D). (18)

Note that this bound differs from Eq. (16) by a log(N) factor. Additionally, although the bound on tsi in Eq. (16) may allow further tightening, the bound on tms in Eq. (18) cannot be generically improved for 0 ⩽ α < D. To see this, we consider a long-range Ising Hamiltonian H=ijJijσizσjz, with Jij=1/rijα. For simplicity, we consider the subset S to be a single site indexed by i and construct operators A=σi+ and B=jiσj+ that are supported on S and S¯, respectively. We can analytically calculate the expectation value of [A(t), B] in an initial state |ψ=12[|0N2+|1N] [54]:

ψ|[A(t),B]|ψ=sin(2tjiJij). (19)

Using Jij=1/rijα, we find that the signaling time of this protocol is t = O(Nα/D−1) for 0 ⩽ α < D, which saturates the many-site signaling-time bound in Eq. (18) [56]. This does not, however, imply that the corresponding scrambling-time bound is tight. In fact, previous work suggests that fast scramblers in all-to-all interacting systems (α = 0) can scramble in time tsc=O(log(N)/N) [36,57]. This suggests that future improvements to the scrambling-time bound may be possible.

Conclusions and outlook.

In this Rapid Communication, we make several advances in bounding the signaling and scrambling times in strongly long-range interacting systems. Our results suggest a number of possible future directions. One is to find the optimal signaling-time bound for general strongly long-range interacting systems. Previously, this has been an outstanding challenge; we now know of a free-particle bound that is tight for α ∈ [0, D/2] and a general bound that is nearly tight as αD. The search for the optimal bound for α ∈ [0, D] has thus been narrowed down significantly. Another direction is to investigate how interactions can speed up signaling. We expect weakly interacting systems to possess a similar signaling-time bound to our free-particle bound, as the dynamics in such systems can often be treated using spin-wave analysis [58]. But for strongly interacting systems, it remains unclear how much speedup one can obtain over noninteracting systems.

Additionally, our bound for signaling to an extensive number of sites hints at a strategy for achieving a better scrambling bound. In particular, the protocol that saturates our many-site signaling bound relies on an initial entangled state, whereas the definition of scrambling assumes that the system begins in a product state. It may be possible to improve the scrambling-time bound by explicitly restricting our attention, when bounding tms, to initial product states.

Finally, we expect that the improved Lieb-Robinson-type bounds derived in this work may lead to a better understanding of the spreading of out-of-time-order correlators [59], the growth of entanglement entropy [54], and thermalization timescales [60] in strongly long-range interacting systems.

In addition, there are connections between Lieb-Robinson-type bounds and the critical scaling of the defects appearing in a quantum system driven across its quantum critical point [the celebrated Kibble-Zurek (KZ) mechanism]. It remains an open question whether the KZ hypothesis can be shown to hold for strongly long-range systems.

Supplementary Material

1

Acknowledgments.

We thank B. Swingle, M. Foss-Feig, J. Garrison, Z. Eldredge, A. Deshpande, and A. Lucas for helpful discussions. M.C.T., A.Y.G., and A.V.G. acknowledge funding by the DoE ASCR Quantum Testbed Pathfinder program (Award No. DE-SC0019040), DoE BES Materials and Chemical Sciences Research for Quantum Information Science program (Award No. DE-SC0019449), NSF PFCQC program, ARO MURI, AFOSR, DoE ASCR Accelerated Research in Quantum Computing program (Award No. DE-SC0020312), AFOSR MURI, ARL CDQI, and NSF PFC at JQI. A.M.C. acknowledges funding by the Army Research Office (MURI Award No. W911NF-16-1-0349), the Canadian Institute for Advanced Research, the National Science Foundation (under Grant No. CCF-1813814), and the U.S. Department of Energy, Office of Science, Office of Advanced Scientific Computing Research, Quantum Algorithms Teams and Quantum Testbed Pathfinder (Award No. DE-SC0019040) programs. Z.X.G. acknowledges support from the NSF RAISE-TAQS program under Grant No. CCF-1839232. A.Y.G. is supported by the NSF Graduate Research Fellowship Program under Grant No. DGE-1840340. M.C.T. is supported in part by the NSF Grant No. NSF PHY-1748958 and the Heising-Simons Foundation.

References

  • [1].Lieb EH and Robinson DW, Commun. Math. Phys 28, 251 (1972). [Google Scholar]
  • [2].Hastings MB, J. Stat. Mech.: Theory Exp (2007) P08024. [Google Scholar]
  • [3].Bravyi S, Hastings MB, and Michalakis S, J. Math. Phys 51, 093512 (2010). [Google Scholar]
  • [4].Saffman M, Walker TG, and Mølmer K, Rev. Mod. Phys 82, 2313 (2010). [Google Scholar]
  • [5].Britton JW, Sawyer BC, Keith AC, Wang CCJ, Freericks JK, Uys H, Biercuk MJ, and Bollinger JJ, Nature (London) 484, 489 (2012). [DOI] [PubMed] [Google Scholar]
  • [6].Yan B, Moses SA, Gadway B, Covey JP, Hazzard KRA, Rey AM, Jin DS, and Ye J, Nature (London) 501, 521 (2013). [DOI] [PubMed] [Google Scholar]
  • [7].Yao NY, Jiang L, Gorshkov AV, Maurer PC, Giedke G, Cirac JI, and Lukin MD, Nat. Commun 3, 800 (2012). [DOI] [PubMed] [Google Scholar]
  • [8].Douglas JS, Habibian H, Hung C-L, Gorshkov AV, Kimble HJ, and Chang DE, Nat. Photonics 9, 326 (2015). [Google Scholar]
  • [9].Gong Z-X, Foss-Feig M, Michalakis S, and Gorshkov AV, Phys. Rev. Lett 113, 030602 (2014). [DOI] [PubMed] [Google Scholar]
  • [10].Foss-Feig M, Gong Z-X, Clark CW, and Gorshkov AV, Phys. Rev. Lett 114, 157201 (2015). [DOI] [PubMed] [Google Scholar]
  • [11].Tran MC, Guo AY, Su Y, Garrison JR, Eldredge Z, Foss-Feig M, Childs AM, and Gorshkov AV, Phys. Rev. X 9, 031006 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Hastings MB and Koma T, Commun. Math. Phys 265, 781 (2006). [Google Scholar]
  • [13].Kastner M, Phys. Rev. Lett 106, 130601 (2011). [DOI] [PubMed] [Google Scholar]
  • [14].Kastner M, Central Eur. J. Phys 10, 637 (2012). [Google Scholar]
  • [15].Storch D-M, van den Worm M, and Kastner M, New J. Phys 17, 063021 (2015). [Google Scholar]
  • [16].Kastner M, Stat J. Mech: Theory Exp (2017) 014003. [Google Scholar]
  • [17].Dauxois T, Ruffo S, Arimondo E, and Wilkens M, Dynamics and Thermodynamics of Systems with Long-Range Interactions: An Introduction (Springer, Berlin/Heidelberg, 2002), pp. 1–19. [Google Scholar]
  • [18].Islam R, Senko C, Campbell WC, Korenblit S, Smith J, Lee A, Edwards EE, Wang C-CJ, Freericks JK, and Monroe C, Science 340, 583 (2013). [DOI] [PubMed] [Google Scholar]
  • [19].Blatt R and Roos CF, Nat. Phys 8, 277 (2012). [Google Scholar]
  • [20].Choi S, Choi J, Landig R, Kucsko G, Zhou H, Isoya J, Jelezko F, Onoda S, Sumiya H, Khemani V, von Keyserlingk C, Yao NY, Demler E, and Lukin MD, Nature (London) 543, 221 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Foss-Feig M, Gong Z-X, Gorshkov AV, and Clark CW, arXiv:1612.07805. [Google Scholar]
  • [22].Defenu N, Enss T, Kastner M, and Morigi G, Phys. Rev. Lett 121, 240403 (2018). [DOI] [PubMed] [Google Scholar]
  • [23].Acevedo OL, Quiroga L, Rodríguez FJ, and Johnson NF, Phys. Rev. Lett 112, 030403 (2014). [DOI] [PubMed] [Google Scholar]
  • [24].Sachdev S and Ye J, Phys. Rev. Lett 70, 3339 (1993). [DOI] [PubMed] [Google Scholar]
  • [25].Gärttner M, Bohnet JG, Safavi-Naini A, Wall ML, Bollinger JJ, and Rey AM, Nat. Phys 13, 781 (2017). [Google Scholar]
  • [26].Kitaev A, KITP Strings Seminar and Entanglement 2015 Program, Feb. 12, April 7, and May 27, 2015, http://online.kitp.ucsb.edu/online/entangled15/kitaev/.
  • [27].Maldacena J, Shenker SH, and Stanford D, J. High Energy Phys 08 (2016) 106. [Google Scholar]
  • [28].Sekino Y and Susskind L, J. High Energy Phys 10 (2008) 065. [Google Scholar]
  • [29].Bentsen G, Gu Y, and Lucas A, Proc. Natl. Acad. Sci. USA 116, 6689 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Chen C-F and Lucas A, arXiv:1905.03682. [Google Scholar]
  • [31]. We will define the quantities tsi and tsc rigorously later in the Rapid Communication.
  • [32].Eisert J, van den Worm M, Manmana SR, and Kastner M, Phys. Rev. Lett 111, 260401 (2013). [DOI] [PubMed] [Google Scholar]
  • [33].Hauke P and Tagliacozzo L, Phys. Rev. Lett 111, 207202 (2013). [DOI] [PubMed] [Google Scholar]
  • [34].Eldredge Z, Gong Z-X, Young JT, Moosavian AH, Foss-Feig M, and Gorshkov AV, Phys. Rev. Lett 119, 170503 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Santos LF, Borgonovi F, and Celardo GL, Phys. Rev. Lett 116, 250402 (2016). [DOI] [PubMed] [Google Scholar]
  • [36].Lashkari N, Stanford D, Hastings M, Osborne T, and Hayden P, J. High Energy Phys. 04 (2013) 022. [Google Scholar]
  • [37]. While the original bound in Ref. [15] was stated in terms of n-body interactions, here we cite the bound on the interaction time as it applies to the specific case of two-body interactions.
  • [38]. The “big-O” notation f (x) = O(g(x)) indicates the existence of constants c1 and N1 such that 0 ⩽ f(x) ⩽ c1g(x) for all x ⩾ N1. We also use the “big-Ω” notation f (x) = Ω(g(x)) to indicate that g(x) = O(f (x)). Finally, we use the “big-Θ” notation f (x) = Θ(g(x)) when both f (x) = O(g(x)) and f (x) = Ω(g(x)). [Google Scholar]
  • [39].Nikolopoulos GM and Jex I, Quantum State Transfer and Network Engineering (Springer, Berlin/Heidelberg, 2014). [Google Scholar]
  • [40].Epstein JM and Whaley KB, Phys. Rev. A 95, 042314 (2017). [Google Scholar]
  • [41]. The strength of the on-site hop Jii is defined this way for technical reasons that we explain in the Supplemental Material [42].
  • [42]. See Supplemental Material at. http://link.aps.org/supplemental/10.1103/PhysRevA.102.010401. for a detailed proof of Eq. (15). The Supplemental Material also shows that Eq. (15) is (in a broad sense) the best bound one can get from Eq. (13).
  • [43].Roberts DA and Swingle B, Phys. Rev. Lett 117, 091602 (2016). [DOI] [PubMed] [Google Scholar]
  • [44].Hayden P and Preskill J, J. High Energy Phys 09 (2007) 120. [Google Scholar]
  • [45].Xu S and Swingle B, Phys. Rev. X 9, 031048 (2019). [Google Scholar]
  • [46].Shenker SH and Stanford D, J. High Energy Phys 03 (2014) 067. [Google Scholar]
  • [47].Hosur P, Qi X-L, Roberts DA, and Yoshida B, J. High Energy Phys 02 (2016) 004. [Google Scholar]
  • [48].Nahum A, Vijay S, and Haah J, Phys. Rev. X 8, 021014 (2018). [Google Scholar]
  • [49].von Keyserlingk CW, Rakovszky T, Pollmann F, and Sondhi SL, Phys. Rev. X 8, 021013 (2018). [Google Scholar]
  • [50].Roberts DA, Stanford D, and Streicher A, J. High Energy Phys 06 (2018) 122. [Google Scholar]
  • [51].Gu Y, Qi X-L, and Stanford D, J. High Energy Phys 05 (2017) 125. [Google Scholar]
  • [52].Pappalardi S, Russomanno A, Žunkovič B, Iemini F, Silva A, and Fazio R, Phys. Rev. B 98, 134303 (2018). [Google Scholar]
  • [53].Chen X and Zhou T, Phys. Rev. B 100, 064305 (2019). [Google Scholar]
  • [54].Gong Z-X, Foss-Feig M, Brandão FGSL, and Gorshkov AV, Phys. Rev. Lett 119, 050501 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Cotler J, Han X, Qi X-L, and Yang Z, J. High Energy Phys 07 (2019) 042. [Google Scholar]
  • [56]. A different protocol for many-site signaling was given in [32]. That result yields a many-site signaling time of tms = O(Nα/D−½), which does not saturate the bound in Eq. (18).
  • [57]. We note that the example given in [36] of a fast scrambler that scrambles completely in time tsc = O(log(N)/N) does not strictly obey the normalization condition hij(t)1/rijα. This does not, however, weaken the claim that there is gap between the scrambling-time bound and the scrambling times of fast-scrambling systems.
  • [58].Cevolani L, Despres J, Carleo G, Tagliacozzo L, and Sanchez-Palencia L, Phys. Rev. B 98, 024302 (2018). [Google Scholar]
  • [59].Luitz DJ and Bar Lev Y, Phys. Rev. A 99, 010105(R) (2019). [Google Scholar]
  • [60].Nandkishore R and Huse DA, Annu. Rev. Condens. Matter Phys 6, 15 (2015). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1

RESOURCES