arxiv:1603.03053v2 [cond-mat.mes-hall] 1 may 2016 · universal chiral quasi-steady states in...

11
Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner, 1 Erez Berg, 2 and Mark S. Rudner 3 1 Physics Department, Technion, 320003 Haifa, Israel 2 Department of Condensed Matter Physics, The Weizmann Institute of Science, Rehovot, 76100, Israel 3 Niels Bohr International Academy and Center for Quantum Devices, University of Copenhagen, 2100 Copenhagen, Denmark (Dated: May 3, 2016) We investigate many-body dynamics in a one-dimensional interacting periodically driven system, based on a partially-filled version of Thouless’s topologically quantized adiabatic pump. The cor- responding single particle Floquet bands are chiral, with the Floquet spectrum realizing nontrivial cycles around the quasienergy Brillouin zone. For generic filling, with either bosons or fermions, the system is gapless; here the driving cannot be adiabatic and the system is expected to rapidly absorb energy from the driving field. We identify parameter regimes where scattering between Flo- quet bands of opposite chirality is exponentially suppressed, opening a long time window where the many-body dynamics separately conserves the occupations of the two chiral bands. Within this intermediate time regime we predict that the system reaches a chiral quasi-steady state. This state is universal in the sense that the current it carries is determined solely by the density of particles in each band and the topological winding numbers of the Floquet bands. This remarkable behav- ior, which holds for both bosons and fermions, may be readily studied experimentally in recently developed cold atom systems. I. INTRODUCTION Topological transport has garnered great attention in condensed matter physics, ever since the discovery of the quantized Hall effects [1, 2]. Traditionally, robust fea- tures of transport have been linked to topological proper- ties of the ground states of many-body systems. Recently, a new paradigm has emerged for altering the topologi- cal properties of band structures using external driving [3–23]. These ideas have sparked a variety of propos- als and initial experiments aimed at realizing so-called Floquet topological insulators in a range of solid state, atomic, and optical systems [24–28]. While the prospect of dynamically controlling band topology is exciting, es- tablishing how and to what extent topological phenom- ena can be observed in such systems raises many crucial and fundamental questions about many-body dynamics in periodically driven systems. The long-time behavior of periodically-driven many- body systems is a fascinating question of current inves- tigation. Several recent theoretical works suggest that closed, interacting, driven many-body systems generi- cally absorb energy indefinitely from the driving field, tending to infinite-temperature-like states in the long time limit [29–31]. In such a state, all correlations are trivial, indicating in particular that any topological fea- tures of the underlying Floquet spectrum are expected to be washed out. On the other hand, several interest- ing exceptions to the infinite temperature fate have been proposed [32–34]. Particularly, heating may be circum- vented by local conservation laws, e.g., in integrable [35] or many-body localized systems [36–43]. A system’s featureless fate at long times may not pre- clude it from exhibiting topological phenomena tran- siently, on timescales shorter than that of its unbounded heating. Indeed, under some conditions such as high fre- a) b) 0 0 ⇡/a 2⇡/a Quasienergy Crystal momentum c) d) -V (t) V (t) J (t) J 0 (t) δJ δV 0 0 500 1000 0.5 1.0 Time (driving periods) Current (1/T ) 0 0 0.5 1.0 Time (driving periods) Current (1/T ) 40 20 T 2T FIG. 1. Equilibration and quasi-steady current in a partially- filled Thouless pump. a) Tight binding model, see Eq. (1). The dimerized hopping amplitudes and on-site potentials are changed adiabatically as depicted by the magenta curve. The origin δJ = δV = 0 is a degeneracy point. b) The Floquet spectrum for ω =0.2J0 and λ = 1, see text below Eq. (1). The right (left) moving Floquet band is colored green (yel- low). c) The period-averaged current vs. time obtained from a numerical simulation of a system of length L = 16 unit cells with N = 8 particles. The initial value of the current depends on the initial state, but after a few driving periods the system reaches a quasi-steady state featuring a current J≈ ρ/T , where ρ = N/L is the density of particles. d) At long times, particle scattering from the right to the left mov- ing band leads to a decay of the current. The parameters used for panels (c) and (d) are ω =0.33J0, λ =0.66 and U =2J0. quency driving [44–47], exponentially-slow heating rates provide long time windows in which interesting “prether- mal” behavior may be observed [47–55]. In this paper we show that a non-trivial long lived quasi-steady state can be stabilized in an interacting arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016

Upload: others

Post on 31-May-2020

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

Universal chiral quasi-steady states in periodically driven many-body systems

Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

1Physics Department, Technion, 320003 Haifa, Israel2Department of Condensed Matter Physics, The Weizmann Institute of Science, Rehovot, 76100, Israel

3Niels Bohr International Academy and Center for Quantum Devices,University of Copenhagen, 2100 Copenhagen, Denmark

(Dated: May 3, 2016)

We investigate many-body dynamics in a one-dimensional interacting periodically driven system,based on a partially-filled version of Thouless’s topologically quantized adiabatic pump. The cor-responding single particle Floquet bands are chiral, with the Floquet spectrum realizing nontrivialcycles around the quasienergy Brillouin zone. For generic filling, with either bosons or fermions,the system is gapless; here the driving cannot be adiabatic and the system is expected to rapidlyabsorb energy from the driving field. We identify parameter regimes where scattering between Flo-quet bands of opposite chirality is exponentially suppressed, opening a long time window where themany-body dynamics separately conserves the occupations of the two chiral bands. Within thisintermediate time regime we predict that the system reaches a chiral quasi-steady state. This stateis universal in the sense that the current it carries is determined solely by the density of particlesin each band and the topological winding numbers of the Floquet bands. This remarkable behav-ior, which holds for both bosons and fermions, may be readily studied experimentally in recentlydeveloped cold atom systems.

I. INTRODUCTION

Topological transport has garnered great attention incondensed matter physics, ever since the discovery of thequantized Hall effects [1, 2]. Traditionally, robust fea-tures of transport have been linked to topological proper-ties of the ground states of many-body systems. Recently,a new paradigm has emerged for altering the topologi-cal properties of band structures using external driving[3–23]. These ideas have sparked a variety of propos-als and initial experiments aimed at realizing so-calledFloquet topological insulators in a range of solid state,atomic, and optical systems [24–28]. While the prospectof dynamically controlling band topology is exciting, es-tablishing how and to what extent topological phenom-ena can be observed in such systems raises many crucialand fundamental questions about many-body dynamicsin periodically driven systems.

The long-time behavior of periodically-driven many-body systems is a fascinating question of current inves-tigation. Several recent theoretical works suggest thatclosed, interacting, driven many-body systems generi-cally absorb energy indefinitely from the driving field,tending to infinite-temperature-like states in the longtime limit [29–31]. In such a state, all correlations aretrivial, indicating in particular that any topological fea-tures of the underlying Floquet spectrum are expectedto be washed out. On the other hand, several interest-ing exceptions to the infinite temperature fate have beenproposed [32–34]. Particularly, heating may be circum-vented by local conservation laws, e.g., in integrable [35]or many-body localized systems [36–43].

A system’s featureless fate at long times may not pre-clude it from exhibiting topological phenomena tran-siently, on timescales shorter than that of its unboundedheating. Indeed, under some conditions such as high fre-

a) b)

00

/a 2/a

Quas

iener

gy

Crystal momentum

c) d)

V (t)V (t)

J(t) J 0(t)

J

V

00

500 1000

0.5

1.0

Time (driving periods)C

urr

ent

(1/T

)0

0

0.5

1.0

Time (driving periods)

Curr

ent

(1/T

)

4020

T

2

T

FIG. 1. Equilibration and quasi-steady current in a partially-filled Thouless pump. a) Tight binding model, see Eq. (1).The dimerized hopping amplitudes and on-site potentials arechanged adiabatically as depicted by the magenta curve. Theorigin δJ = δV = 0 is a degeneracy point. b) The Floquetspectrum for ω = 0.2J0 and λ = 1, see text below Eq. (1).The right (left) moving Floquet band is colored green (yel-low). c) The period-averaged current vs. time obtained froma numerical simulation of a system of length L = 16 unitcells with N = 8 particles. The initial value of the currentdepends on the initial state, but after a few driving periodsthe system reaches a quasi-steady state featuring a currentJ ≈ ρ/T , where ρ = N/L is the density of particles. d) Atlong times, particle scattering from the right to the left mov-ing band leads to a decay of the current. The parameters usedfor panels (c) and (d) are ω = 0.33J0, λ = 0.66 and U = 2J0.

quency driving [44–47], exponentially-slow heating ratesprovide long time windows in which interesting “prether-mal” behavior may be observed [47–55].

In this paper we show that a non-trivial long livedquasi-steady state can be stabilized in an interacting

arX

iv:1

603.

0305

3v2

[co

nd-m

at.m

es-h

all]

1 M

ay 2

016

Page 2: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

2

periodically driven system, in the low frequency limit.We focus our attention to one dimension (1d), whereThouless showed that cyclic adiabatic driving in filled-band/gapped many-body systems leads to topologically-quantized charge pumping [56]; this phenomenon was re-cently observed in cold atoms experiments [57, 58]. Anexample of a time-dependent tight-binding model withthis property is illustrated in Fig. 1a [see Eq. (1) be-low]. In the adiabatic limit, the system’s single-particleFloquet spectrum exhibits a non-trivial winding of eachband [5]: the quasi energy changes by ω ≡ 2π/T (whereT is the driving period) as the crystal momentum changesfrom 0 to 2π/a, where a is the lattice constant [5, 59].A characteristic Floquet spectrum with such winding (or“chiral”) bands is shown in Fig. 1b.

The average current carried by the system over onedriving period is quantized when one of the Floquetbands is completely filled with fermions, and the otheris completely empty. What happens when the bands areinitially only partially filled, or if the system is comprisedof bosons? Here the system lacks a many-body gap,and the evolution cannot be considered to be adiabatic.In the absence of interactions any value of the pumpedcharge per cycle is possible, depending on the details ofthe Hamiltonian and the initial state. However, in theinteracting case we identify a parameter regime whereheating naturally drives the system to a quasi-steadystate with universal properties, reflecting the topologi-cal nature of the Floquet band structure.

We study dynamics in the situation where one of thechiral Floquet bands is initially partially filled, while theother is empty. We assume that the minimum gap be-tween the two bands of the single particle Hamiltonian(minimized over all times and quasi-momenta), ∆, ismuch larger than max[~ω,U,W ], where U and W arethe interaction strength and the maximum bandwidth,respectively (see below for a precise definition of W ).Based on an analysis of low and high order scatteringrates, we argue that that (i) regardless of details, thesystem equilibrates on a short time scale τintra to a chi-ral infinite-temperature-like state in which all momentumstates in one of the chiral bands are equally populated,while the other band remains nearly empty; (ii) in thisstate, the average pumped charge per period is approx-imately wρ, where w is the winding number of the par-tially filled band (see below) and ρ is the density of par-ticles [60]; and (iii) the quasi-steady state persists for along time scale τinter, that can be larger than τintra bymany orders of magnitude. At times t τintra, the sys-tem relaxes to a state in which the values of all localobservables are as in an infinite temperature state. Inparticular, the current tends to zero. This physical pic-ture is supported by exact numerical simulations of finitesystems, see Figs. 1c,d.

Model. – For concreteness, we consider a 1d lattice withtwo sites per unit cell, labeled A and B. The hopping ma-trix elements and on-site potentials are time dependent(see Fig. 1a). The lattice is populated by a finite density

ρ of identical fermions or bosons per unit cell.We write the Hamiltonian as H(t) = H0(t) + Hint,

where the single particle part H0(t) is given by

H0(t) = −J(t)∑j

c†j,Acj,B − J ′(t)∑j

c†j,B cj+1,A + h.c.

+ V (t)∑j

(c†j,Acj,A − c†j,B cj,B). (1)

Here, c†j,A (c†j,B) creates a particle of type A (B) in unitcell j. To avoid later ambiguity, we use hats to de-note creation and annihilation operators throughout thiswork. Below we take J(t) = J0+δJ(t), J ′(t) = J0−δJ(t),and V (t) = V0 + δV (t), with δJ(t) = λJ0 cosωt andδV (t) = V1 sinωt. For simplicity we fix V1 = 3λJ0 in allsimulations presented below.

We consider a local interaction

Hint = U∑j

nj(nj − 1), nj = c†j,Acj,A + c†j,B cj,B . (2)

The intra-unit-cell form of the interaction in Eq. (2) isconvenient for the analysis below. However, we do notexpect our conclusions to depend on this choice.

The single particle Floquet spectrum corresponding toH0(t) is found by seeking solutions to the Schrodingerequation which satisfy [61]: |Ψ1P(t)〉 = e−iε1Pt|Φ1P(t)〉,with |Φ1P(t + T )〉 = |Φ1P(t)〉. We decompose the peri-odic function |Φ1P(t)〉 in terms of an infinite set of (non-

normalized) discrete Fourier modes |ϕ(m)1P 〉:

|Ψ1P(t)〉 = e−iε1Pt∑m

|ϕ(m)1P 〉e−imωt. (3)

The full time-dependent evolution of |Ψ1P(t)〉 is specifiedby the quasienergy ε1P and a vector of Fourier coefficients

ϕ1P =

...

|ϕ(−1)1P 〉|ϕ(0)

1P 〉|ϕ(1)

1P 〉...

. (4)

Throughout this work we choose the quasi-energies of allsingle particle states to lie within a fundamental Floquet-Brillouin zone, 0 ≤ ε1P < 2π/T .

The Fourier coefficients comprising |Ψ(t)〉 are deter-mined by an eigenvalue equation εϕ = H0ϕ, wherethe “extended Hamiltonian” H0 is constructed from theFourier decomposition of H0(t) (see Appendix A) andϕ without a ket symbol stands for a column vector ofFourier modes as in Eq. (4). We use calligraphic sym-bols for matrices in the space of Fourier coefficients. Forharmonic driving, H0(t) = Hdc+Λeiωt+Λ†e−iωt, the ma-trix H takes a simple block tri-diagonal form in harmonic(m) space: (H0)mm′ = (Hdc + mω)δmm′ + (Λδm,m′−1 +Λ†δm,m′+1). The single particle Floquet spectrum is

Page 3: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

3

shown in Fig. 1b for parameters specified in the caption.For each of the two bands, we assign a winding numberw, such that the quasi-energy band winds by wω whenthe quasi-momentum k changes from 0 to 2π/a. In ourcase, the two bands have winding numbers w = +1 andw = −1, and we refer to them as the right-moving (R)and left-moving (L) bands, respectively. More generally,a non-trivial winding is achieved in the adiabatic limitwhen the curve (δJ(t), δV (t)) encircles the origin (as inFig. 1a).

II. MANY-BODY DYNAMICS

We now turn to the many-body dynamics of this sys-tem. We consider the situation where the system is ini-tialized with a finite density of particles in the net right-moving (R) Floquet band, shown in green in Fig. 1b. Theinitial momenta of the particles are arbitrary.

To investigate the timescales for intra-band and inter-band scattering, we develop a perturbative analysis ofthe many-body dynamics of the system. The perturba-tion series is organized in terms of powers of the interac-tion strength U . Crucially, we work in a Floquet picturewhere the time-dependent driving is first taken into ac-count exactly, to all orders in the driving. As above, wework in the extended space of Fourier coefficients, wherethe many-body Floquet eigenstates are described by theeigenvectors of the extended Hamiltonian, H = H0 + U ,with Umm′ = Hintδmm′ .

The extended Hamiltonian defines a (static) eigen-value problem that yields the Fourier coefficients de-scribing Floquet eigenstates. One may also use the ex-tended Hamiltonian to generate an effective evolution inthe extended space, in an auxiliary time variable τ , viai∂τϕ(τ) = Hϕ(τ). For the stroboscopic times τ = nT ,where n is an integer and T is the driving period of theoriginal problem, the “evolved” vector of Fourier coeffi-cients ϕ(nT ) precisely captures the state of the systemin the physical Hilbert space at the corresponding timet = nT (see Appendix B). Using this mapping we obtaintransition rates for the stroboscopic evolution by employ-ing standard Green’s function techniques to the auxiliaryevolution problem in the extended space.

Our aim is to calculate the rate at which particlesare scattered into the left-moving (L) band. For weakinteractions and short times, it is natural to view thisprocess in terms of a perturbation series in the inter-action U . We express the auxiliary-time evolution ofthe Fourier vector ϕ(τ) in terms of its Fourier trans-form, ϕ(τ) =

∫∞−∞ dτ eiΩτ ϕ(Ω). In terms of the extended

Green’s function G0(Ω) = (Ω−H0 + iδ)−1 and T-matrixT (Ω) = U + UG0(Ω)T (Ω), we have

ϕ(Ω) = i [G0(Ω) + G0(Ω)T (Ω)G0(Ω)]ϕ0, (5)

where ϕ0 is the Fourier vector corresponding to the “free”initial state in which all particles are initialized in single-particle Floquet eigenstates in the right-moving (R) band

of the non-interacting system (see Appendix C for detailsof the construction of ϕ0).

A. Born approximation

As a first step, we investigate the scattering ratesto leading order in U , i.e., in the Born approximationT (Ω) ≈ U . This approximation captures the leading-order behavior of two-particle scattering, which one mayexpect to be relevant for weak interactions and low den-sities (see discussion below and Refs. [62–64]).

Within the Born approximation, the transition rate

is given by Γ ≈ 2π∑f 6=0 δ(Ef − E0)|ϕ†f Uϕ0|2. This is

Fermi’s golden rule, adapted for a Floquet system [65].Here f labels all final “free” Floquet eigenstates, andEf are their eigenenergies (with respect to the ex-tended Hamiltonian H0). We break Γ into two parts,Γ = Γintra+Γinter, corresponding to intra-band and inter-band scattering, respectively. The latter scattering pro-cesses transfer one or more particles from the R to the Lband.

Figures 2a,b show the two-particle scattering rates fora pair of bosons in the R band, with momenta k1, k2,to scatter either to states within the R band (processeswe denote as RR→RR, Fig. 2a), or to states with oneparticle in the R band and one in the L band (RR→RL,Fig. 2b). The two-particle scattering rates are given by

Γ2P

L=∑f

1

|∆vf ||ϕ†2P,f Uϕ2P,0|2, (6)

where ϕ2P,0 and ϕ2P,f are the wavefunctions of the initialand final two-particle Floquet states (see Appendix C),∆vf is the difference of group velocities of the two out-going particles, and the summation is over all final statesthat satisfy quasi-energy and quasi-momentum conser-vation. The factor 1/|∆vf | comes from the density ofoutgoing states.

Our key observation is that the inter-band scatteringrates are strongly suppressed compared with the intra-band ones. For the parameters chosen, the mean RR →RL scattering rate is down by a factor ∼ 10−12 comparedto the average RR → RR rate. The average RR → LLrate (not shown) is suppressed by a factor of ∼ 10−10

relative to RR → RL. The bright lines visible in the in-terband scattering rate, Fig. 2b, are associated with sin-gle particle resonances; such resonances yield significantinterband scattering rates only in exponentially small re-gions of phase space (see Appendix A).

The suppression of inter-band relative to intra-bandscattering originates from the matrix element in Eq. (6).Since U is diagonal in Fourier harmonics, the matrix el-ement for inter-band scattering is suppressed if the ini-tial and final states have support in different regions inharmonic space. In the adiabatic limit ω ∆, this is in-deed the case. Figure 2c shows two representative single-particle Floquet wavefunctions with quasi-momenta k1,

Page 4: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

4

0 0.2 0.4 0.6 0.8 1

−4

−3

−2

−1

0

1

2

3

4

t / T

E k(t)

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

-50 -40 -30 -20 -10 0 100

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

a) b)

c) d)k1

0 /a 2/a0

a

2a

k2

RR ! RR

0

k1

0 /a 2/a0

a

2a

k2

RR ! RL

-10

-14

2

0Floquet harmonic, m

-20-400

0.2

0

0.2

initial

final0.4

0.4 RR ! RL

RR ! RR

Ek(t

)

tTT/20

W

0

Floquetharmonic, m

-12

-84

-2-16

Inte

nsi

ty,|'

(m)

2P

|2

|'(m)1P |2

log10

2P

log10

2P

−15

−10

−5

05

10

15

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9 1

10

10

FIG. 2. Two-particle scattering in a partially-filled Thoulesspump. a) Intraband scattering rate Γ2P = Γ2P/(J0L), eval-uated for bosons within Fermi’s golden rule, Eq. (6), withforward scattering contribution removed. Here k1 and k2 arethe momenta of two incident particles, and we take λ = 0.56,ω = 0.2J0, and U = 0.67J0. b) Same as above, for interbandscattering in which one particle is scattered from the right tothe left moving band. c) Single particle Floquet states. Onthe left, the shaded region spans the energies of the instan-taneous bands as a function of time. The significant Fourier

components |ϕ(m)1P |

2 = |〈ϕ(m)1P |ϕ

(m)1P 〉|

2 of a Floquet state αwith momentum k fall between the minimum and maximumvalues of the instantaneous energies Eα,k(t). d) Fourier com-

ponents |ϕ(m)2P |

2 = |〈ϕ(m)2P |ϕ

(m)2P 〉|

2 of incoming and outgoingtwo particle Floquet states. For intraband scattering (bot-tom) the Fourier components overlap. For interband scat-tering (top), the overlap is strongly suppressed leading to asuppression of the interband scattering rate.

k2, one from each band. The support of each state in har-monic space corresponds to the energy window spannedby the instantaneous energy, Eα,k(t) (the eigenvalue ofH0(t) for band α = R,L), with 0 ≤ t < T [66]. Thisenergy window is bounded by W , which we define asW = maxk,tER,k(t)−mink,tER,k(t), see Fig. 2c. Outsideof this window, the Floquet wavefunctions decay rapidly.The separation of the Floquet states of the two bands inharmonic space can be derived by mapping the Floquetproblem to a Zener tunneling problem in a weak electricfield (see Appendix A).

Figure 2d shows representative two-particle states thatparticipate in either inter- or intra-band scattering. Thetwo-particle states are constructed as convolutions of twosingle-particle states. For intra-band scattering (RR →RR), the initial and final states occupy the same regionin harmonic space. In contrast, for inter-band scattering(RR → RL) the initial and final states are separated inharmonic space by a gap of order ∆/ω. (Note that thisrequires the energy spread of the single particle states, oforder W , to be smaller than ∆.) Hence, at least withinthe Born approximation, inter-band scattering is strongly

suppressed with respect to intra-band scattering.

B. Higher order contributions

Next, we consider higher order contributions to theinter-band scattering rate [67]. Importantly, the strongsuppression of the Born-level interband scattering ratearises from the exponentially small overlap of harmonic-space wave functions of the initial and final states, alongwith the fact that the time-independent interaction con-serves the harmonic index. As we now show, this expo-nential factor may be avoided at higher orders in per-turbation theory, trading the small matrix element forhigher powers of the interaction, U . Optimizing overthe competition between small U and small overlaps, wefind an optimal order Nmin that dominates the scatter-ing amplitude. In this way we argue that the scatteringrate should be a power law in U , with an exponentialsuppression in the inverse of the driving frequency, 1/ω.

To analyze the scattering amplitudes at higher ordersin the interaction, we return to the expansion of the T-matrix,

T (Ω) = U + UG0(Ω)U + UG0(Ω)UG0(Ω)U + · · · . (7)

Recall that the operators T (Ω), G0(Ω) and U are de-fined in the “extended space” of Fourier harmonics. Aswe will see, the additional state-space dimensions of theextended space play an important role in describing en-ergy (photon) absorption from the drive.

To facilitate the T-matrix analysis, we define a ba-sis of non-interacting eigenstates of the extended spaceHamiltonian H0. For each non-interacting N parti-cle Floquet state we associate a label ξ = kn, αn,n = 1 . . . N , denoting the quasimomenta and the bandindices αn = R,L of all particles. As in the singleparticle case, cf. Eq. (4), the N particle non-interactingFloquet state |Ψξ(t)〉 is represented by a vector of Fouriercoefficients which we denote by ϕξ. We use the conven-tion that the many-body quasi-energy is given by the sumof single particle quasi-energies, εξ =

∑n ε1P,knαn , with

0 ≤ ε1P,knαn < 2π/T as above. Under this convention,the quasi-energies εξ generically fall outside of the fun-damental Floquet zone. The detailed construction of themany-body Fourier vectors ϕξ is given in Appendix C.

In the extended space representation this spectrum isrepeated over and over again, shifted by integer multi-ples of ω, with the corresponding eigenvectors likewiseshifted in harmonic space. Therefore, with each Floquetstate ϕξ defined above, we can associate a whole family

of eigenstates ϕξ,ν with components ϕ(m)ξ,ν = ϕ

(m+ν)ξ

and quasienergies εξ + νω. (Note that complete infor-mation about the physical Floquet spectrum is capturedby eigenstates with a single ν.) In terms of these “copy

Page 5: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

5

a)

b)

c)Floquet harmonic, m

...

... n!

m!

n!

m!

Ener

gyden

om

inat

or

1 . . . n n + 1 . . . Step index

Virtual statesRR

RR RL

RL

Floquet harmonic, m

m

|'(m

)2P,

|2|'

(m)

2P,

|2Born approximation

Nmin-order scattering

Nmin 1

FIG. 3. Matrix elements and energy denominators for high-order scattering. a) In the Born approximation, the RR→ RLscattering matrix element is (super) exponentially suppressedby the small overlap of harmonic space wave functions. b)At an order Nmin ∼ ∆/(δmω) in perturbation theory, a non-suppressed matrix element between initial and final states forthe RR → RL process is constructed by moving through asequence of off-shell intermediate states, shifted from one an-other by δν . δm harmonics. c) Energy denominators forthe virtual states in the process described in b). Multiplyingthese energy denominators gives the expression in Eq. (10).

states” we write the extended Green’s function G0(Ω) as

G0(Ω) =∑ξ,ν

ϕξ,νϕ†ξ,ν

Ω− εξ − νω + iδ. (8)

Crucially, the “copy states” defined above play an im-portant role as off-shell virtual intermediate states in theperturbation theory.

At each order, the scattering amplitude involves aproduct of many-body matrix elements of the form

ϕ†ξ,νUϕξ′,ν′ . Such a matrix element can be expressedin the time-domain as

ϕ†ξ,νUϕξ′,ν′ =1

T

∫ T

0

dt e−i(ν−ν′)ωt〈ϕξ(t)|Hint|ϕξ′(t)〉,

(9)

where |ϕξ(t)〉 =∑m e−imωt|ϕ(m)

ξ 〉 is the periodic part of

the (non-interacting) many-body Floquet wave function.Importantly, Hint is a two-body operator. There-

fore, computation of the matrix element in Eq. (9) canbe reduced to the problem of evaluating matrix ele-

ments between two-particle Floquet states, M(ν−ν′)ξ1ξ2,ξ′1ξ

′2

=

1T

∫ T0dt e−i(ν−ν

′)ωt〈ϕ2P, ξ1ξ2(t)|Hint|ϕ2P, ξ′1ξ′2(t)〉. By fur-

ther understanding the properties of these two-particlematrix elements, we may thus characterize the variousterms in the high-order perturbation theory.

For simplicity we focus on the case W ω ∆,where the spread δm of the two particle Floquet wave-

functions in Fourier space is of order δm = O(1), seeFig. 2c,d, and Fig. 3. The dependence of the two-particlematrix elements on ν−ν′ is distinguished by the Floquetband indices of the incoming and outgoing states. Inparticular, as seen for the special case of the on-shellprocess depicted in Fig. 2d, for an RR→ RR (intra-band) transition the harmonic-space wave functions givean order-1 contribution (no suppression) to Eq. (9) for|ν − ν′| . δm. For an RR→ RL matrix element, order-1overlap of the harmonic-space wave functions is achievedfor ∆

ω − δm . |ν − ν′| . ∆ω + δm. Note that this non-

suppressed interband scattering matrix element necessar-ily describes an off-shell process. We neglect “doubleinterband” processes RR → LL, the rates of which areheavily suppressed compared with those of the primaryRR → RL decay process.

Using the rules above, we see that at high order in per-turbation theory it is possible to construct an interbandscattering amplitude that avoids the (super) exponen-tial suppression arising from the tiny overlap of harmonicspace wave functions that appears in the Born approxi-mation. At each step in the perturbation theory, U mayconnect virtual states with |ν − ν′| ∼ δm. Therefore, asindicated in Fig. 3b, there is an order Nmin ∼ ∆

δmω atwhich the initial (RR) and final (RL) two-particle statescan be connected with no suppression due the harmonicspace wave functions.

Viewing the construction of the Nmin-order scatteringamplitude as a sequence of steps, each application of Ureduces the harmonic-space separation between the vir-tual intermediate and the final (on-shell) state. Thesesuccessive off-shell intermediates bring energy denomi-nators of larger and larger magnitudes. After n steps,one particle is transferred from the R to the L band,giving a jump in the sequence of energy denominators.The terms in this sequence, in units of δmω, are of order−1,−2, . . . ,−n,Nmin−n,Nmin−n−1, . . . , 1, see Fig. 3c.We write the corresponding transition amplitude as

A(n)inter ≈

(−1)nann!(Nmin − n)!(δmω)Nmin

. (10)

Here, an ∝ UNmin contains a sum of matrix elementsof the interaction between the intermediate states. Thisgives the following rough estimate for the inter-band scat-tering rate (see Appendix D for details):

Γinter ∝∣∣∣∑n

A(n)inter

∣∣∣2 ∼ (αU

) 2∆δmω

. (11)

Here, α is an O(1) numerical constant.Physically, in the regime W ω ∆ that we have

considered, the dominant inter-band scattering processesare associated with Nmin scattering events, each one in-volving the absorption of δm energy quanta of ω from thedriving field, such that the total absorbed energy is ∆.We expect such processes to dominate as long as ω is nottoo small compared to W . If W ω, the typical change

Page 6: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

6

in m in every scattering event scales with the width of thesingle-particle Floquet wavefunctions in harmonic space,δm ∼W/ω. Hence we would have Nmin ∼ ∆

W in Eq. (11),and for low enough frequencies we expect Γinter to satu-rate.

III. NUMERICAL SIMULATIONS

To further study the many-body dynamics we haveperformed exact numerical simulations of the dynamicsof finite-size systems. To minimize the Hilbert space di-mension we consider fermions in these simulations. Webelieve that the qualitative behavior does not depend onthe quantum statistics, but leave the detailed study ofthe bosonic case to future investigations.

In each simulation, we initialize the system in a Slaterdeterminant state where N momentum states in theright-moving (R) Floquet band are occupied, and theleft-moving (L) band is empty. The results do not de-pend sensitively on which states in the R band are ini-tially occupied, except at very short times. The particlesmove on a lattice of L unit cells (with two sites each),with periodic boundary conditions. The largest systemwe studied contained 8 particles with L = 16 unit cells.

Figures 1c,d show the period-averaged current,

J (nT ) =

∫ (nT+1)T

nTT

dt′1

L

∑j

J ′(t′)⟨ic†j,Bcj+1,A + h.c.

⟩t′,

(12)as a function of nT , the number of periods elapsed (seefigure caption for model parameters). At time t = 0,the system carries a current that depends on the ini-tial state. Over a timescale τintra ∼ 10T the current re-laxes to a value J ≈ ρ/T , see Fig. 1c (in this simulation,ρ = 0.5). We ascribe this time scale to relaxation withinthe R band, while the L band remains almost unpopu-lated. Examining the occupation numbers of momentumstates confirms this interpretation (see Appendix E); fort & τintra, the occupation numbers in the R band areapproximately uniform and all close to ρ. The averagegroup velocity of states in the R band is vg = a/T . There-fore, the average current J = ρ

avg observed in Fig. 1c isas expected for a uniform particle distribution in the rightmoving band.

At longer times the current undergoes a much slowerdecay process toward zero, with a time scale τinter of sev-eral hundred periods (Fig. 1d). During this process thepopulation of the L band gradually increases. For t→∞the occupation numbers of all the momentum states inboth bands tend toward ρ/2, corresponding to a maxi-mally randomized infinite-temperature-like state.

For times up to several times τintra, we find that the av-erage (total) occupation number in the L band increasesapproximately linearly with time. The slope of the lineargrowth, which we define as Γinter ≡ 1/τinter, is found tobe only weakly dependent on system size for L between10 and 16 (see Appendix E).

-3

-4

-5

-6

-7

-8-1.2 -0.8 -0.4

/!log10[U/]

log10[

inte

r/(J

0L

)]

!/J0

0.200.28

0.40

5

4 8 12/!

U/

4 8 12 16

0.580.73

0.881

slop

e

FIG. 4. Interband excitation rate Γinter obtained from exactnumerical evolution of a system with 8 fermions on 32 sites(L = 16 unit cells). Left panel: Log-log plot showing de-pendence of Γinter/(J0L) on interaction strength U , for threefixed values of frequency ω and for band structure parameterλ = 0.56. Trend lines are linear fits to the data, confirming apower law dependence consistent with Eq. (11). Inset: Powerlaw exponent as a function of ω, extracted from linear fits tothe data. Right panel: Log-linear plot showing the drivingfrequency dependence of Γinter/(J0L), indicating an exponen-tial dependence on 1/ω, for the same model parameters.

Figure 4 presents the rate Γinter as a function of modelparameters. The dependence of Γinter on U for differentvalues of ω (with all other parameters fixed) is presentedin the left panel of Fig. 4 in a log-log plot, showing apower-law dependence on U . The power depends lin-early on 1/ω (left panel, inset). The right panel showsthe dependence on ω at fixed U . Clearly, Γinter scalesexponentially with 1/ω. For the lowest frequencies westudied, corresponding to ∆/ω ≈ 16 (where ∆ is theminimum instantaneous gap), the rate reaches ∼ 10−7

in units of J0, indicating a very long-lived quasi-steadystate where only the R band is populated. We also findthat the behaviour of Γinter shown in Fig 4 is not sen-sitive to the details of the band structure, and persiststhroughout a wide range of values of the band structureparameter λ.

Interestingly, the numerically obtained Γinter(U, ω) isconsistent with the form of Eq. (11). This suggests that,within our model, the inter-band relaxation process isdominated by “multi-photon assisted” scattering events,where many energy quanta are absorbed from the drivingfield.

IV. DISCUSSION

In this work we studied the dynamics of a periodically-driven many-body system where the driving frequencyis much smaller than the instantaneous inter-band gapthroughout the driving period. When the system is pre-pared such that one of the bands is initially empty, whilethe other is partially occupied, a very long intermediatetime window emerges, in which a universal quasi-steadystate is realized. The quasi-steady state carries a robustcurrent whose value depends solely on the density of par-

Page 7: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

7

ticles and the topological winding number of the under-lying single-particle Floquet spectrum. In this sense, thecombination of periodic driving and interactions leads toa non-equilibrium “prethermalized” state with topologi-cal properties.

Beyond the demonstration that such non trivial long-lived quasi-steady states are possible in slowly drivenmany-body systems, our results may be directly appli-cable to recent experiments on cold atomic fermions [58]and bosons [57], where quantized pumping has recentlybeen demonstrated. For the parameter regime statedabove, initializing the system with a fractional filling ofone of the bands should result in a current carrying quasi-steady state.

The mechanism presented here is expected to applyto any many-body periodically driven system where thedriving frequency is much smaller than the instantaneousgap to some subset of excitations. Under such condi-tions, the high-energy excitations can remain “frozen”for a long intermediate time window, leading to a quasi-steady state with non-trivial properties. This work opensmany intriguing theoretical and experimental questionsabout many-body dynamics and topology in driven sys-tems, which will be interesting to explore in future work.

ACKNOWLEDGMENTS

We thank Dima Abanin, Michael Knap, and AnatoliPolkovnikov for illuminating discussions, and David Co-

hen for technical support. NL and EB acknowledge fi-nancial support from the European Research Council(ERC) under the European Unions Horizon 2020 researchand innovation programme (grant agreement No 639172).MR gratefully acknowledges the support of the VillumFoundation and the People Programme (Marie Curie Ac-tions) of the European Unions Seventh Framework Pro-gramme (FP7/2007-2013) under REA grant agreementPIIF-GA-2013-627838.

Appendix A: Single particle Floquet wavefunctions

In this Appendix we elaborate on the extended zoneformalism, and apply it to the single-particle Floquetstates. In particular, we derive the (localized) form ofthe Floquet wavefunctions in harmonic space by relatingthe problem to Zener tunneling in a two-band system ina linear potential.

Consider a single-particle Hamiltonian of the formH0(t) = Hdc + Λeiωt + Λ†e−iωt. (The same formalismapplies to any time-periodic Hamiltonian.) We assumethat the single particle Hamiltonian is diagonal in mo-mentum space, and focus on the Floquet states |Ψk(t)〉for a single value of the crystal momentum k. For sim-plicity we also assume that the system has only twobands. The corresponding 2 × 2 Bloch Hamiltonian is

H0,k(t) = Hdc,k + Λkeiωt + Λ†ke

−iωt.

Inserting a Floquet state of the form |Ψk(t)〉 = e−iεkt∑m |ϕ

(m)k 〉e−imωt into the time-dependent Schrodinger equa-

tion, we get that the quasi-energy εk and the Fourier components |ϕ(m)k 〉 satisfy:

. . . ΛkΛ†k Hdc,k + ω Λk

Λ†k Hdc,k ΛkΛ†k Hdc,k − ω Λk

Λ†k. . .

...

|ϕ(−1)k 〉|ϕ(0)k 〉|ϕ(1)k 〉...

= εk

...

|ϕ(−1)k 〉|ϕ(0)k 〉|ϕ(1)k 〉...

. (A1)

The operator on the left hand side of this equation is theFloquet “extended zone” operator, or “extended Hamil-tonian,” H0,k. Note that H0.k has a block-tridiagonalform, analogous to that of a nearest-neighbor tight bind-ing Hamiltonian. In this picture, the terms proportionalto ω on the diagonal correspond to a linear potential onthe m-lattice.

In the adiabatic limit, ω ∆, we first solve the prob-lem with ω = 0. The diagonal “linear potential” is laterintroduced as a perturbation. For ω = 0, the problem istranslationally invariant in harmonic (m) space; we cansolve it by Fourier transforming along the m-direction.This amounts to transforming from frequency back to

time domain (where t plays the role of “momentum” forthe m-space tight-binding problem). The secular equa-tion then becomes

H0(t)|Ψα(t)〉 = Eα(t)|Ψα(t)〉, (A2)

which is nothing but the Schrodinger equation for theinstantaneous eigenstates and eigenenergies. Here, α =R,L is the band index. The two bands are separated bya gap, |EL,k(t)− ER,k(t)| ≥ ∆.

Next, we consider the effect of the linear potentialterm. The situation is illustrated in Fig. 5a. The problemis equivalent to the well-known Zener tunneling problemof a two-band semiconductor in an electric field. The

Page 8: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

8

W

m

"m!

a)

−60 −40 −20 0 20 40−35

−30

−25

−20

−15

−10

−5

0

Fits: Right: −35.5*| (x+1) / 30 | 1.5, left: −25*| (x+11) / 30 | 1.5

m

log10(|

(m)

1P

|2 )

0

10

20

30

b)

A|m m0|32

0 202040

FIG. 5. Effective Zener tunneling problem for the Floquetstates in harmonic space. a) The Floquet problem, Eq. (A1),is equivalent to a tight-binding problem in harmonic (m)space, with a linear potential term mω. The effective bandshave a characteristic width W and gap ∆. The gray regionsindicate the “classically allowed” regions for each fixed valueof “energy,” as determined by the sum of the band energiesand the linear potential. The Floquet wavefunctions in thetwo bands (illustrated by the red and blue bars) are stronglylocalized in the classically allowed regions, and decay rapidlybeyond these regions. b) Modulus squared of a representativeFloquet wave function as a function of the Fourier index m.Sufficiently far from the maximum, the wavefunction decaysas exp(−A|m−m0|3/2), where A and m0 are constants.

states in both bands become localized in the “classicallyallowed region” in harmonic space, whose characteristicsize is of order Wα/ω (where Wα = maxk,tEαk(t) −mink,tEαk(t) is an effective bandwidth).

Within this picture, states in the R and L bands thatare close in quasienergy (within ∼ ω of each other) areseparated in harmonic space by a spacing of the order of∆/ω. The tails of the wavefunctions decay rapidly intothe classically forbidden region; in the limit ω Wα,there is a broad region where the wavefunctions decay as

e−A|m−m0|3/2

(see Fig. 5b), where A is a band structuredependent dimensionless constant and m0 is the bound-ary of the classically forbidden region. (This form is ex-pected from a Wentzel–Kramers–Brillouin treatment ofthe problem, where m is treated as a continuous vari-able). At asymptotically long distances, larger than sev-eral times Wα/ω, the form of the wavefunction crossesover to e−B|m−m0| log |m−m0|, where B is another dimen-sionless constant.

The tunneling matrix element between the two bandsis hence strongly suppressed in the adiabatic limit. Thehybridization between the two bands is expected to bevery small, unless their energies are very close to eachother. As k varies, the bands nearly cross at a set of kpoints (see Fig. 1b in the main text); at these crossingpoints the two bands hybridize, and there is an avoidedcrossing gap whose magnitude is exponentially small inthe adiabatic limit. For a generic band structure, thehybridization between the bands is significant only withinexponentially small regions in k space around the near-crossing points. These hybridizations are responsible forthe bright lines appearing in Fig. 2b in the main text.

Appendix B: Stroboscopic dynamics using theextended zone Hamiltonian

The extended zone Hamiltonian H in harmonic space(shown in Eq. (A1) for the single-particle case) is de-signed such that its spectrum and eigenstates correspondto the quasi-energies and Floquet states, respectively. Itcan also be used to generate the stroboscopic dynamicsat times t = nT , where n is an integer, starting from anarbitrary initial state.

In order to see this, consider a Floquet state solu-tion of the time-dependent Schrodinger equation gen-erated by the physical Hamiltonian H(t), of the form|ψ(t)〉 = e−iεt

∑m |ϕ(m)〉e−imωt. Compare this state to

a solution of the auxiliary-time Schrodinger equation inharmonic (extended) space, generated by H (see maintext):

ϕext(τ) = e−iετ (. . . , |ϕ(−1)〉, |ϕ(0)〉, |ϕ(1)〉, . . .)T . (B1)

(We use Dirac bra-ket notation for states in the physicalHilbert space, whereas vectors in the extended space arewritten without Dirac notation.) We relate the “evolved”state in the extended space to a state in the physi-cal Hilbert space at stroboscopic times t = τ = NTby summing over the harmonic components of ϕext(τ):|ϕext(t = NT )〉 = e−iεNT

∑m |ϕ(m)〉. At these strobo-

scopic times, we find that |ψ(t = nT )〉 and |ϕext(t = nT )〉coincide.

The reasoning above can be extended to the time evo-lution of an arbitrary initial state, |ψ0〉. This is done byexpanding |ψ0〉 in terms of Floquet eigenstates, consider-ing the time evolution with respect to either H(t) or H,and comparing the time-evolved wavefunctions at timest = nT .

Appendix C: Construction of non-interacting(“free”) many-body Floquet states

We build up the non-interacting N -particle state ϕξ

one harmonic at a time. First, we define a single particleFloquet state ϕ1P,ξi , as in Eq. (4), for each ξi = ki, αi,with i = 1 . . . N . As used throughout the main text, thesingle particle quasienergies ε1P,ξi are taken in the inter-val 0 ≤ ε1P,ξi < 2π/T . Within this convention we define

a set of creation operators φ(m′)†ξ , for all m′, using the

identification |ϕ(m′)1P,ξ 〉 = φ

(m′)†ξ |0〉. Our analysis holds for

both bosons and fermions, where the creation and an-nilation operators respectively satisfy the commutation

(+) and anticommutation (−) relations [φ(m)ξ , φ

(m′)ξ′ ]± =

[φ(m)†ξ , φ

(m′)†ξ′ ]± = 0 and [φ

(m)ξ , φ

(m′)†ξ′ ]± = 〈ϕ(m)

1P,ξ|ϕ(m′)1P,ξ′〉.

The m-th Fourier component of the many-body stateϕξ is determined by a convolution over the single particleharmonics:

|ϕ(m)ξ 〉 =

∑mi

δm,∑imi

[N∏i=1

φ(mi)†ξi

]|0〉. (C1)

Page 9: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

9

0 10 20 30 40

Time (driving periods)

0

0.2

0.4

0.6

0.8

1

N(R

)

k

0 10 20 30 40

Time (driving periods)

0

0.005

0.01

0.015

0.02

N(L

)

k

FIG. 6. Particle distribution in single particle Floquet states

in the right and left moving Floquet bands, N(R)k and N

(L)k , as

a function of time. In both figures, we indicated only the max-

imal and minimal occupations, maxkN(α)k and minkN

(α)k , by

shading the area between them. In both plots the yellow line

indicates the average occupation 1L

∑kN

(α)k within the given

band. As can be seen from the left figure, the distribution inthe left moving band relaxes to an approximately uniform dis-tribution within a short time scale of a few driving periods.The right figure clearly shows the interband scattering ratefrom the right to the left moving Floquet band. The param-eters for this figure were ω = 0.27J0, λ = 0.56, U = 1.67J0,L = 16, and ρ = 0.5.

In the above, the sum extends over all sets of N integersmi, for i = 1, .., N . Note that the quasi-energy of the

many-body state ϕξ is εξ =∑Ni=1 ε1P,ξi , with the con-

vention 0 ≤ ε1P,ξi < 2π/T specified above. Thus εξ isuniquely specified, and generically falls outside the fun-damental Floquet-Brillouin zone.

Appendix D: Estimate of the interband scatteringrate

Here we describe the steps leading to the estimate ofthe interband scattering rate [Eq. (6) of the main text].We begin with the term in the expression for the ampli-tude of order Nmin, in which a particle is scattered fromthe L to the R band after n steps, with 0 ≤ n ≤ Nmin−1.This amplitude (in the limit W ω) is given by

A(n)inter ≈

(−1)nann!(Nmin − n)!(δmω)Nmin

. (D1)

Using Stirling’s formula for n,Nmin 1, we may replacen!(Nmin − n)! ≈ eNmin[log(Nmin)−f(n/Nmin)], where f(x) =1 + x log(x) + (1 − x) log(1 − x). The function f(x) isbounded and f(x) = O(1) for 0 ≤ x ≤ 1. Therefore

we approximate the factor e−Nminf(n/Nmin) by β−Nmin1 to

leading exponential accuracy in the limit of large Nmin,where β1 = O(1) is some constant.

In order to get a rough estimate of the interband scat-tering rate Γinter, we must examine the factor an inEq. (6), which contains a sequence of matrix elementsof the interaction term U between the initial, interme-diate, and final states. The dominant dependence of anon Nmin is expected to be of the form an ∼ (β2U)Nmin ,

where β2 = O(1). Hence, summing the amplitudes A(n)inter

over n, taking the modulus of the square, and usingNmin ∼ ∆

δmω , we get to the estimate for Γinter quotedin Eq. (6) of the main text. While this estimate is ratherrough, our numerical results for Γinter show excellentagreement with the form predicted in Eq. (6).

Appendix E: Further details of the numericalsimulations

In our numerical simulations, the many-body wave-function |Ψ(t)〉 is time-evolved numerically, using theHamiltonian H(t) [Eqs. (1,2)], for up to 1000 driving pe-riods. The simulations are performed using a finite timestep, ∆t, and using a Trotter-Suzuki decomposition ofthe evolution operator within each time step. Most of thesimulations were done with ∆t = T/500. For ω < 0.2J0,we used ∆t = T/850. We have verified the results do notchange upon decreasing ∆t further, even for the longesttimes simulated.

To extract the inter-band rates plotted in Fig. 3 ofthe main text, we computed the distribution of par-ticles in the different single-particle Floquet states attimes which correspond to integer multiples of the driv-

ing period. We thus calculate N(α)k (t = mT ) =

〈Ψ(mT )|ψ†k,αψk,α|Ψ(mT )〉 where ψ†k,α is the creation op-

erator for the Floquet state |ψ(0)k,α〉 with momentum k,and the index α =R, L indicates the right or left movingFloquet band.

Typical particle distributions in the Floquet bandsare plotted in Fig. 6. States within the right movingband quickly become nearly equally populated, withprobabilities close to 0.5 (corresponding to the densityρ = N/L = 0.5, taken in the simulation). This indicatesthe establishment of a quasi-infinite-temperature state,restricted to the right moving (R) band. After aninitial transient of a few driving periods, the popu-lation in the left moving (L) band increases linearlywith time, with a small rate, while the populationin the right moving band decreases with the samerate. The rates shown in Fig. 3 of the main text wereobtained by considering the rate of increase of theaverage population in the left moving band (slope ofthe yellow line in the right panel of Fig. 6), Γinter =1L

∑k

[N

(L)k (t = mT )−N (L)

k (t = m0T )]/ [(m−m0)T ]

with m = 20 and m0 = 5.The largest system that we could reach with our nu-

merical simulations included 8 particles on 32 sites (i.e.,L = 16 unit cells and density ρ = 0.5). To verify thatthe rates reported in Fig. 3 of the main text do not suf-fer from substantial finite size effects, we studied the sizedependence of the interband scattering rate, normalizedto the length of the system, Γinter/(J0L), while keepingthe density fixed. For the parameter regimes plotted inFig. 3 of the main text, we found that the size-normalizedrate is only weakly size dependent between L = 10 and

Page 10: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

10

FIG. 7. Size dependence of the interband scattering rate nor-malized by the size of the system, Γ/(J0L). As the number ofunit cells, L, is changed, the density of particles is held fixedat ρ = 0.5. The parameters used for this figure were ω = 0.3,λ = 0.56, U = 2.

L = 16, and appears to be saturating by L = 16. Thisindicates that at L = 16 finite size effects are small anddo not change the results qualitatively. A representativeplot of the finite size dependence of the size-normalizedrate can be found in Fig. 7.

[1] K. v. Klitzing, G. Dorda, and M. Pepper, Phys. Rev.Lett. 45, 494 (1980).

[2] R. B. Laughlin, Phys. Rev. B 23, 5632 (1981).[3] T. Oka and H. Aoki, Phys. Rev. B 79, 081406 (2009).[4] J.-i. Inoue and A. Tanaka, Phys. Rev. Lett. 105, 017401

(2010).[5] T. Kitagawa, E. Berg, M. Rudner, and E. Demler, Phys.

Rev. B 82, 235114 (2010).[6] N. H. Lindner, G. Refael, and V. Galitski, Nat. Phys. 7,

490 (2011).[7] N. H. Lindner, D. L. Bergman, G. Refael, and V. Galitski,

Phys. Rev. B 87, 235131 (2013).[8] Z. Gu, H. A. Fertig, D. P. Arovas, and A. Auerbach,

Phys. Rev. Lett. 107, 216601 (2011).[9] T. Kitagawa, T. Oka, A. Brataas, L. Fu, and E. Demler,

Phys. Rev. B 84, 235108 (2011).[10] P. Delplace, A. Gomez-Leon, and G. Platero, Phys. Rev.

B 88, 245422 (2013).[11] Y. T. Katan and D. Podolsky, Phys. Rev. Lett. 110,

016802 (2013).[12] D. E. Liu, A. Levchenko, and H. U. Baranger, Phys. Rev.

Lett. 111, 047002 (2013).[13] P. Titum, N. H. Lindner, M. C. Rechtsman, and G. Re-

fael, Phys. Rev. Lett. 114, 056801 (2015).[14] G. Usaj, P. M. Perez-Piskunow, L. E. F. Foa Torres, and

C. A. Balseiro, Phys. Rev. B 90, 115423 (2014).[15] L. E. F. Foa Torres, P. M. Perez-Piskunow, C. A. Bal-

seiro, and G. Usaj, Phys. Rev. Lett. 113, 266801 (2014).[16] L. D’Alessio and M. Rigol, arXiv:1409.6319 (2014).[17] H. Dehghani, T. Oka, and A. Mitra, Phys. Rev. B 90,

195429 (2014).[18] H. Dehghani, T. Oka, and A. Mitra, arXiv:1412.8469

(2014).[19] M. Sentef, M. Claassen, A. Kemper, B. Moritz, T. Oka,

and T. Freericks, J.Kand Devereaux, Nat. Comm. 6, 7047(2015).

[20] K. I. Seetharam, C.-E. Bardyn, N. H. Lindner, M. S.Rudner, and G. Refael, Phys. Rev. X 5, 041050 (2015).

[21] T. Iadecola, T. Neupert, and C. Chamon, Phys. Rev. B

91, 235133 (2015).[22] J. Klinovaja, P. Stano, and D. Loss, arXiv:1510.03640

(2015).[23] Y. Gannot, arXiv:1512.04190 (2015).[24] Y. H. Wang, H. Steinberg, P. Jarillo-Herrero, and

N. Gedik, Science 342, 453 (2013).[25] A. Eckardt and E. Anisimovas, New Journal of Physics

17, 093039 (2015).[26] G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat,

T. Uehlinger, D. Greif, and T. Esslinger, Nature 515,237 (2014).

[27] M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer,D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Sza-meit, Nature 496, 196 (2013).

[28] W. Hu, J. C. Pillay, K. Wu, M. Pasek, P. P. Shum, andY. D. Chong, Phys. Rev. X 5, 011012 (2015).

[29] A. Lazarides, A. Das, and R. Moessner, Phys. Rev. E 90,012110 (2014).

[30] L. DAlessio and M. Rigol, Phys. Rev. X 4, 041048 (2014).[31] A. Lazarides, A. Das, and R. Moessner, Phys. Rev. Lett.

115, 030402 (2015).[32] R. Citro, E. G. D. Torre, L. DAlessio, A. Polkovnikov,

M. Babadi, T. Oka, and E. Demler, Annals of Physics360, 694 (2015).

[33] A. Chandran and S. L. Sondhi, arXiv:1506.08836 (2015).[34] I. Kukuljan and T. Prosen, arXiv:1512.06601 (2015).[35] A. Lazarides, A. Das, and R. Moessner, Phys. Rev. Lett.

112, 150401 (2014).[36] P. Ponte, A. Chandran, Z. Papic, and D. A. Abanin,

Annals of Physics 353, 196 (2014).[37] D. Abanin, W. De Roeck, and F. Huveneers,

arXiv:1412.2752 (2014).[38] V. Khemani, A. Lazarides, R. Moessner, and S. L.

Sondhi, arXiv:1508.03344 (2015).[39] C. W. von Keyserlingk and S. L. Sondhi,

arXiv:1602.02157 (2016).[40] C. W. von Keyserlingk and S. L. Sondhi,

arXiv:1602.06949 (2016).[41] A. C. Potter, T. Morimoto, and A. Vishwanath,

Page 11: arXiv:1603.03053v2 [cond-mat.mes-hall] 1 May 2016 · Universal chiral quasi-steady states in periodically driven many-body systems Netanel H. Lindner,1 Erez Berg,2 and Mark S. Rudner3

11

arXiv:1602.05194 (2016).[42] D. V. Else and C. Nayak, arXiv:1602.04804 (2016).[43] R. Roy and F. Harper, arXiv:1602.08089 (2016).[44] D. A. Abanin, W. De Roeck, and F. m. c. Huveneers,

Phys. Rev. Lett. 115, 256803 (2015).[45] D. A. Abanin, W. De Roeck, and W. W. Ho,

arXiv:1510.03405 (2015).[46] A. Eckardt and E. Anisimovas, New Journal of Physics

17, 093039 (2015).[47] M. Bukov, M. Heyl, D. A. Huse, and A. Polkovnikov,

arXiv:1512.02119 (2015).[48] J. Berges, S. Borsanyi, and C. Wetterich, Phys. Rev.

Lett. 93, 142002 (2004).[49] M. Eckstein, M. Kollar, and P. Werner, Phys. Rev. Lett.

p. 056403 (2009).[50] M. Moeckel and S. Kehrein, New Journal of Physics 12,

055016 (2010).[51] L. Mathey and A. Polkovnikov, Phys. Rev. A 81, 033605

(2010).[52] M. Bukov, S. Gopalakrishnan, M. Knap, and E. Demler,

Phys. Rev. Lett. 115, 205301 (2015).[53] E. Canovi, M. Kollar, and M. Eckstein,

arXiv:1507.00991v3 (2016).[54] T. Kuwahara, T. Mori, and K. Saito, Annals of Physics

367, 96 (2016).[55] T. Mori, T. Kuwahara, and K. Saito, arXiv:1509.03968v2

(2015).[56] D. J. Thouless, Phys. Rev. B 27, 6083 (1983).[57] M. Lohse, C. Schweizer, O. Zilberberg, M. Aidelsburger,

and I. Bloch, arXiv:1507.02225; Nature Physics, in press(2015).

[58] S. Nakajima, T. Tomita, S. Taie, T. Ichinose,H. Ozawa, L. Wang, M. Troyer, and Y. Takahashi,arXiv:1507.02223; Nature Physics, in press (2016).

[59] For any finite drive frequency ω, minigaps open at thecrossing between the Floquet bands. These minigaps aresuppressed exponentially in 1/ω.

[60] A conservative estimate of the corrections to this valueof the current is O[(U/∆)2], due to virtual transitionsbetween the bands.

[61] We take ~ = 1 throughout.[62] S. Choudhury and E. J. Mueller, Phys. Rev. A 91, 023624

(2015).[63] T. Bilitewski and N. R. Cooper, Physical Review A 91,

033601 (2015).[64] M. Genske and A. Rosch, Phys. Rev. A 92, 062108

(2015).[65] The Fermi’s golden rule rates are evaluated using “free”

initial and final states, as taken at t = 0. For t & τintra,the state locally appears similar to an infinite tempera-ture state (within the R band), which can be describedas a uniform mixture of all free states in the band.

[66] We use a convention where the quasi-energies lie in thewindow [0, ω).

[67] The consideration here are similar to the ones employedin the problem of doublon decay in a Mott insulator. E.g.,see: R, Sensarma, D. Pekker, M. D. Lukin, E. Dem-ler, Phys. Rev. Lett. 103, 035303 (2009); Z. Lenarcic,M. Eckstein, and P. Prelovsek, Phys. Rev. B 92, 201104(2015).