1department of physics, imperial college london, london … · quasi-periodically driven quantum...

10
Quasi-periodically driven quantum systems Albert Verdeny, 1 Joaquim Puig, 2 and Florian Mintert 1 1 Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom 2 Department of Mathematics, Universitat Polit` ecnica de Catalunya, 08007 Barcelona, Spain Floquet theory provides rigorous foundations for the theory of periodically driven quantum sys- tems. In the case of non-periodic driving, however, the situation is not so well understood. Here, we provide a critical review of the theoretical framework developed for quasi-periodically driven quantum systems. Although the theoretical footing is still under development, we argue that quasi- periodically driven quantum systems can be treated with generalizations of Floquet theory in suitable parameter regimes. Moreover, we provide a generalization of the Floquet-Magnus expansion and argue that quasi-periodic driving offers a promising route for quantum simulations. I. INTRODUCTION The dynamics of quantum systems induced by a time- dependent Hamiltonian attracts attention from various communities. Chemical reactions can be controlled with driving induced by laser beams [1], and driving atoms permits to investigate their electronic structure [2]. Suit- ably chosen driving sequences permit to investigate dy- namics in macro-molecular complexes [3] and there exist phases in solid state systems that can be accessed only in the presence of driving [4, 5]. A neat bridge between quantum optics and solid state physics is built by the fact that periodically driven atomic gases can be employed as quantum simulators for models of solid state theory [6, 7]. Solving the Schr¨ odinger equation with a time- dependent Hamiltonian calls for different mathematical techniques than those applied in situations with time- independent Hamiltonians. Differential equations with time-dependent coefficients have been investigated thor- oughly and, in particular, developments regarding re- ducibility are appreciated, since they permit to un- derstand driven systems in terms of time-independent Hamiltonians [8]. The foundation for this is laid by the Floquet theorem [9], which relates a periodically time-dependent with a constant Hamiltonian. This mathematical theorem pro- vides the basis for experiments that employ periodically driven quantum systems for quantum simulations of sys- tems with time-independent Hamiltonians. Such exper- iments have led to the experimental observation of e.g. coherent suppression of tunneling [1012], spin-orbit cou- pling [13, 14], synthetic magnetism [1517], ferromag- netic domains [18], or topological band structures [19, 20] The specific time-dependence of the driving force plays a crucial role in the dynamics that driven systems can undergo. Yet, despite the possibility to experimentally tune it, very simple driving protocols are usually em- ployed, which can significantly limit the performance of the simulations [21] and restrict the range of accessible dynamics [22, 23]. In this context, pulse-shaping techniques have been in- troduced in order to achieve the simulation of the desired effective dynamics in an optimal fashion [21, 23]. Yet, the restriction to periodic driving is a limitation, and quasi- periodic driving, i.e. driving with a time-dependence characterized by several frequencies that can be irra- tionally related, promises substantially enhanced control over the quantum system at hand. Since the use of quasi- periodic driving [2426], however, implies that Floquet theorem is not applicable, the mathematical foundation is far less solid than in the case of periodic driving. Generalizations of Floquet’s theorem to quasi-periodic driving have been pursued both in the quantum physics/chemistry literature [27, 28] and in the math- ematical literature [8, 2931]. The former perspective approaches quasi-periodic driving as a limiting case of periodic systems, while the mathematical literature ap- proaches quasi-periodicity without resorting to results from periodic systems. Beyond the fundamentally dif- ferent approaches, also the findings in the different com- munities are not always consistent with each other. The goal of the present article is twofold. On the one hand, we discuss prior literature on the generalization of Floquet’s theorem to quasi-periodic systems, and at- tempt to overview over what findings have been verified to mathematical rigour, and what findings are rather based on case studies and still lack a general, rigorous foundation. On the other hand, we aim at studying the possibility to use quasi-periodically driven systems for quantum simulations. We consider quasi-periodic Hamiltonians H(t) that can be defined in terms of a Fourier-like representation of the form H(t)= X n H n e in·ωt , (1) where ω =(ω 1 , ··· d ) is a finite-dimensional vector of real frequencies that are irrationally related, and n = (n 1 , ··· ,n d ) is a vector of integers such that n · ω = n 1 ω 1 + ··· + n d ω d . Moreover, the norm of coefficients H n are considered to decay sufficiently fast with |n|. The main underlying question of the present work is the possibility to express the time-evolution operator U (t) of a quasi-periodically driven system in terms of a generalized Floquet representation of the form U (t) ? = U Q (t)e -iH Q t , (2) where H Q is a time-independent Hermitian operator and U Q (t)= n U n e in·ωt is a quasi-periodic unitary charac- arXiv:1603.03923v1 [quant-ph] 12 Mar 2016

Upload: hoangkhanh

Post on 04-Apr-2018

215 views

Category:

Documents


2 download

TRANSCRIPT

Quasi-periodically driven quantum systems

Albert Verdeny,1 Joaquim Puig,2 and Florian Mintert1

1Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom2Department of Mathematics, Universitat Politecnica de Catalunya, 08007 Barcelona, Spain

Floquet theory provides rigorous foundations for the theory of periodically driven quantum sys-tems. In the case of non-periodic driving, however, the situation is not so well understood. Here,we provide a critical review of the theoretical framework developed for quasi-periodically drivenquantum systems. Although the theoretical footing is still under development, we argue that quasi-periodically driven quantum systems can be treated with generalizations of Floquet theory in suitableparameter regimes. Moreover, we provide a generalization of the Floquet-Magnus expansion andargue that quasi-periodic driving offers a promising route for quantum simulations.

I. INTRODUCTION

The dynamics of quantum systems induced by a time-dependent Hamiltonian attracts attention from variouscommunities. Chemical reactions can be controlled withdriving induced by laser beams [1], and driving atomspermits to investigate their electronic structure [2]. Suit-ably chosen driving sequences permit to investigate dy-namics in macro-molecular complexes [3] and there existphases in solid state systems that can be accessed onlyin the presence of driving [4, 5]. A neat bridge betweenquantum optics and solid state physics is built by the factthat periodically driven atomic gases can be employed asquantum simulators for models of solid state theory [6, 7].

Solving the Schrodinger equation with a time-dependent Hamiltonian calls for different mathematicaltechniques than those applied in situations with time-independent Hamiltonians. Differential equations withtime-dependent coefficients have been investigated thor-oughly and, in particular, developments regarding re-ducibility are appreciated, since they permit to un-derstand driven systems in terms of time-independentHamiltonians [8].

The foundation for this is laid by the Floquet theorem[9], which relates a periodically time-dependent with aconstant Hamiltonian. This mathematical theorem pro-vides the basis for experiments that employ periodicallydriven quantum systems for quantum simulations of sys-tems with time-independent Hamiltonians. Such exper-iments have led to the experimental observation of e.g.coherent suppression of tunneling [10–12], spin-orbit cou-pling [13, 14], synthetic magnetism [15–17], ferromag-netic domains [18], or topological band structures [19, 20]

The specific time-dependence of the driving force playsa crucial role in the dynamics that driven systems canundergo. Yet, despite the possibility to experimentallytune it, very simple driving protocols are usually em-ployed, which can significantly limit the performance ofthe simulations [21] and restrict the range of accessibledynamics [22, 23].

In this context, pulse-shaping techniques have been in-troduced in order to achieve the simulation of the desiredeffective dynamics in an optimal fashion [21, 23]. Yet, therestriction to periodic driving is a limitation, and quasi-

periodic driving, i.e. driving with a time-dependencecharacterized by several frequencies that can be irra-tionally related, promises substantially enhanced controlover the quantum system at hand. Since the use of quasi-periodic driving [24–26], however, implies that Floquettheorem is not applicable, the mathematical foundationis far less solid than in the case of periodic driving.

Generalizations of Floquet’s theorem to quasi-periodicdriving have been pursued both in the quantumphysics/chemistry literature [27, 28] and in the math-ematical literature [8, 29–31]. The former perspectiveapproaches quasi-periodic driving as a limiting case ofperiodic systems, while the mathematical literature ap-proaches quasi-periodicity without resorting to resultsfrom periodic systems. Beyond the fundamentally dif-ferent approaches, also the findings in the different com-munities are not always consistent with each other.

The goal of the present article is twofold. On the onehand, we discuss prior literature on the generalizationof Floquet’s theorem to quasi-periodic systems, and at-tempt to overview over what findings have been verifiedto mathematical rigour, and what findings are ratherbased on case studies and still lack a general, rigorousfoundation. On the other hand, we aim at studying thepossibility to use quasi-periodically driven systems forquantum simulations.

We consider quasi-periodic HamiltoniansH(t) that canbe defined in terms of a Fourier-like representation of theform

H(t) =∑n

Hnein·ωt, (1)

where ω = (ω1, · · · , ωd) is a finite-dimensional vector ofreal frequencies that are irrationally related, and n =(n1, · · · , nd) is a vector of integers such that n · ω =n1ω1 + · · ·+ndωd. Moreover, the norm of coefficients Hn

are considered to decay sufficiently fast with |n|.The main underlying question of the present work is

the possibility to express the time-evolution operatorU(t) of a quasi-periodically driven system in terms ofa generalized Floquet representation of the form

U(t)?= U†Q(t)e−iHQt, (2)

where HQ is a time-independent Hermitian operator andUQ(t) =

∑n Une

in·ωt is a quasi-periodic unitary charac-

arX

iv:1

603.

0392

3v1

[qu

ant-

ph]

12

Mar

201

6

2

terized by the same fundamental frequency vector ω asthe quasi-periodic Hamiltonian H(t). If the frequencyvector ω defining the quasi-periodicity of H(t) containsonly one element, i.e. d = 1, the Hamiltonian becomesperiodic with period 2π/ω1 and the decomposition in Eq.(2) reduces to the usual Floquet representation, which isknown to exist. However, if ω contains more than oneelement, i.e. d > 1, the Hamiltonian is not periodic andthe representation in Eq. (2) is a priori not guaranteed.

The possibility to represent the time-evolution opera-tor of a quasi-periodically driven system as in Eq. (2) isdirectly related to the problem of reducibility [32, 33] offirst-order differential equations with quasi-periodic coef-ficients, which is still an ongoing problem in the mathe-matics community [34]. Unlike their periodic counter-parts, linear differential equations with quasi-periodiccoefficients cannot always be reduced to constant co-efficients by means of a quasi-periodic transformation[35, 36], although a quasi-periodic Floquet reducibilitytheory does exist [29, 37].

Generalizations of Floquet theory to quasi-periodicallydriven systems have been derived also from a lessmathematical perspective. Many-mode Floquet theory(MMFT) [27, 38–40] is based on physical assumptions ofthe underlying time-dependent Hamiltonian, and it hasbeen successfully applied to a variety of systems rang-ing from quantum chemistry [27, 41] to quantum optics[42, 43]. However, it does not seem to have an entirelyrigorous footing yet.

In this article, we address these different perspectivesand argue that, despite gaps in a general mathemati-cal footing, concepts from regular Floquet theory can betranslated directly to quasi-periodically driven systems,especially in fast-driving regimes, i.e. the regime of quan-tum simulations.

In Sec. II we introduce notation and preliminary con-cepts of Floquet theory that will be used throughout thearticle. In Sec. III, we revise critically the MMFT andpoint out aspects of the derivation that cast doubts onthe general validity of the proof. In Sec. IV we arguehow, nevertheless, the general formalism of MMFT canstill lead to valid results, in agreement with prior work[28, 38–40, 42–44]. In Sec. V, we derive a generalizationof the Floquet-Magnus expansion [45], which provides aperturbative exponential expansion of the time-evolutionoperator that has the desired structure. With this, we ad-vocate the possibility to identify effective Hamiltoniansthat characterize well the effective dynamics of quasi-periodically driven systems in a fast-driving regime, andexemplify in Sec. VI the results with a quasi-periodicallydriven Lambda system.

II. FLOQUET THEORY

Floquet’s theorem [9] asserts that the Schrodingerequation

i∂tU(t) = H(t)U(t), (3)

characterizing the time-evolution operator U(t) of a sys-

tem described by a periodic Hamiltonian H(t) = H(t +T ), is reducible. That is, there exists a periodic unitaryUP (t) = UP (t + T ) that transforms the Schrodinger op-

erator K(t) = H(t)− i∂t into

UP (t)K(t)U†P (t) = HF − i∂t , (4)

where HF = UP (t)H(t)U†P (t)−iUP (t)(∂tU†P (t)) is a time-

independent Hamiltonian [46]. As a consequence, thetime-evolution operator of the system can be representedas the product

U(t) = U†P (t)e−iHF t, (5)

with UP (0) = 1. This decomposition is of central impor-tance in the context of quantum simulations with period-ically driven systems because it ensures that, in a suitablefast-driving regime, the dynamics of the driven systemcan be approximated in terms of the time-independentHamiltonian HF [6, 7].

The eigenstates |εk〉 of the Hamiltonian HF form abasis in the system Hilbert space H, on which the pe-riodic Hamiltonian H(t) acts. Thus, any vector |φ(0)〉characterizing the initial state of the system can be writ-ten as a linear combination of the eigenstates |εk〉. Con-sequently, the decomposition of the time-evolution op-erator in Eq. (5) implies that time-dependent states

|φ(t)〉 = U(t)|φ(0)〉 can be expressed as a linear combina-tion with time-independent coefficients of Floquet statesof the form

|φk(t)〉 = e−iεkt|uk(t)〉, (6)

where εk are the eigenvalues of HF (also termed

quasienergies), and |uk(t)〉 = U†P (t)|εk〉 = |uk(t + T )〉are periodic state vectors.

The quasienergies εk play a very important role inthe dynamics of driven systems. They can be calcu-lated after inserting the Floquet states in Eq. (6) into

the Schrodinger equation i∂t|φk(t)〉 = H(t)|φk(t)〉, whichyields

K(t)|uk(t)〉 = εk|uk(t)〉 . (7)

Eq. (7) formally describes an eigenvalue problem resem-bling the time-independent Schrodinger equation, wherethe periodic states |uk(t)〉 play an analogous role of sta-tionary states. Due to the time-dependence of the statesand the action of the derivative in K(t), however, thediagonalization in Eq. (7) cannot be straightforwardlysolved with standard matrix diagonalization techniques.

For this reason, it is often convenient to formulate theproblem in a Fourier space where the operator K(t) istreated as an infinite-dimensional time-independent op-erator [47, 48].

A. Time-independent formalism

State vectors of the driven system are defined on thesystem Hilbert space H, where time is regarded as a pa-

3

rameter. In order to arrive at a formalism in which theparameter ‘time’ does not appear explicitly, one exploitsthe fact that the states |u(t)〉 in H that have a peri-odic time-dependence can be defined on a Floquet Hilbertspace F = H⊗L2(T), where L2(T) is the Hilbert space ofperiodic functions [48]. In this Floquet space, time is notregarded as a parameter but rather as a coordinate of thenew Hilbert space. The explicit time-dependence of thesystem can then be removed by adopting a Fourier rep-resentation of the periodic states in the space F . Fourierseries permit the characterization of periodic functions interms of a sequence of its Fourier coefficients. Formally,this can be described through an isomorphism betweenthe space L2(T) of periodic functions and the space l2(Z)of square-summable sequences. This isomorphism allowsone to map the exponential functions einωt, which form abasis in L2(T), to states |n〉, which define an orthonormalbasis in l2(Z).

In this manner, periodic states |u(t)〉 =∑n |un〉einωt

are mapped to states

|u〉 =∑n

|un〉 ⊗ |n〉, (8)

while periodic operators A(t) =∑nAne

inωt can bemapped to

A =∑n

An ⊗ σn, (9)

were the ladder operators σn =∑m |m + n〉〈m| satisfy

σn|m〉 = |n+m〉. Similarly, the derivative operator −i∂tis mapped to

D = 1⊗ ωn, (10)

with the number operator n =∑n n|n〉〈n| satisfying

n|n〉 = n|n〉 and the commutation relation [n, σm] =mσm.

Consequently, the isomorphism between F and H ⊗l2(Z) permits one to treat the periodic system within aFourier formalism by mapping the Schrodinger operatorK(t) = H(t)− i∂t to the operator [49]

K =∑n

Hn ⊗ σn + 1⊗ ωn . (11)

The time-evolution operator U(t) of the system can thenbe calculated via the relation [47]

U(t) =∑n

〈n|e−iKt|0〉einωt , (12)

which can be readily verified by inserting Eq. (12) intothe Schrodinger equation and using the explicit form ofK given in Eq. (11), as we explicitely demonstrate inAppendix for illustrative purposes.

The operator K in Eq. (11) is often represented as aninfinite-dimensional matrix. The Floquet states in Eq.(6) can be obtained from the diagonalization of K [47].In order to find the Floquet decomposition of the time-evolution operator in Eq. (5), however, it is not necessary

to completely diagonalize the operator K. Instead, theoperator K needs to be brought into the block-diagonalform

KB = UP KU†P = HF ⊗ 1 + 1⊗ ωn (13)

by means of a unitary transformation

UP =∑n

Un ⊗ σn. (14)

The block-diagonalization in Eq. (13) describes the coun-terpart in the present time-independent formalism ofthe transformation in Eq. (4) such that, if the block-diagonalization is achieved, the Floquet HamiltonianHF and the periodic unitary UP (t) =

∑n Une

inωt arestraightforwardly obtained from Eqs. (13) and (14).

III. MANY-MODE FLOQUET THEORYREVISED

The many-mode Floquet theory (MMFT) [27, 38–40]was introduced in the context of quantum chemistry as ageneralization of Floquet theory to treat systems witha quasi-periodic time dependence. The derivation ofMMFT is rooted on Floquet’s theorem and its proposedgenerality contrasts with other results derived with morerigorous approaches. In this section, we revise the deriva-tion of MMFT and challenge aspects of the proof thatquestion its general validity.

The derivation of the MMFT [27] consists in approx-imating the quasi-periodic Hamiltonian H(t) by a pe-riodic Hamiltonian and, then, using Floquet theory todemonstrate the existence of a generalized Floquet de-composition for the time-evolution operator of the sys-tem. The derivation [27] starts by considering a quasi-periodic Hamiltonian H(t) =

∑nHne

in·ωt in Eq. (1)and approximating the different elements ωi of the fre-quency vector ω by a fraction. Then, a small fundamen-tal driving frequency ω is identified such that the differentirrationally-related frequencies ωi are expressed as

ωi ≈ Ni ω, (15)

with some integers Ni. In this manner, the quasi-periodically driven Hamiltonian H(t) can be approxi-mated by the periodic Hamiltonian

H(t) =∑n

Hnein·Nωt, (16)

where N = (N1, · · · , Nd). The validity of the approxi-

mation H(t) ≈ H(t) for a certain time window impor-tantly depends on the good behavior of the Hamiltonianand on the approximation in Eq. (15), which can beperformed with any desired accuracy. When the approx-imation H(t) ≈ H(t) is satisfied with sufficient accuracy,the time-evolution operator U(t) of the quasi-periodicallydriven system can be also approximated by the time-evolution operator U(t) that is induced by the periodic

approximated Hamiltonian H(t), i.e. U(t) ≈ U(t).

4

The next step in the derivation aims at demonstratingthat time-evolution operator U(t) of the periodic Hamil-

tonian H(t) can be approximately represented by a gen-eralized Floquet decomposition analogous to Eq. (2).Specifically, the aim is to express the periodic unitaryUP (t) of the Floquet decomposition in Eq. (5) in termsof a Fourier series of the form

UP (t) =∑n

Unein·Nωt, (17)

which contains only specific Fourier components, since,in general, not all integers can be expressed as n · Nwith a vector of integers n. If this was possible foran arbitrarily small frequency ω, the unitary UP (t) inEq. (17) would approximate a quasi-periodic unitaryUQ(t) =

∑n Une

in·ωt and the time-evolution operatorof the quasi-periodically driven system could be well ap-proximated by the sought generalized Floquet decompo-sition in Eq. (2).

The MMFT derivation [27] considers, for concreteness,a quasi-periodic Hamiltonian H(t) in Eq. (1) with d = 2;that is, the frequency vector contains only two compo-nents: ω1 and ω2. Moreover, the only non-vanishing co-efficientsHn of the quasi-periodic Hamiltonian areH(0,0),H(±1,0), and H(0,±1). Despite this specific choice, how-ever, the possibility to generalize the results to Hamilto-nians containing more frequencies is claimed.

In order to demonstrate the possibility to write theunitary in Eq. (17), the derivation in [27] makes useof the time-independent Floquet formalism described inSec. II. First of all, the operator K in Eq. (11) is defined(using a slightly different notation) for the approximated

periodic Hamiltonian H(t) in Eq. (16) [50]. Then, a

block-diagonal structure is given to K as a first step toachieve the desired structure of the time-evolution oper-ator.

The block-diagonal structure is obtained by ‘rela-belling’ each vector |n〉 that forms a basis of l2(Z) (intro-duced in Sec. II A) as

|np〉, (18)

where the vector of integers n = (n1, n2) and the integerp are found by solving the Diophantine equation

n =

2∑i=1

niNi + p (19)

for all n and for the integers Ni in Eq. (15). Thereafter,a tensor product structure is given to the Hilbert spacel2(Z), such that the state vectors in Eq. (18) are writtenin the tensor product form |np〉 = |n〉|p〉, with |n〉 =|n1〉|n2〉 and where n1, n2, and p can take all integer

values. In this manner, the operator K in Eq. (11) isdescribed to be rewritten in the block-diagonal form

Kbd = K ⊗ 1 + 1⊗ ωp, (20)

with

K =∑n

Hn ⊗ σn + 1⊗ ω · n. (21)

The ladder operator σn and number operators n and p in-troduced in Eq. (21) are defined as σn =

∑m |m+n〉〈m|,

n =∑

n n|n〉〈n|, and p =∑p p|p〉〈p|, respectively, where

the summations include all possible values of n and p.The notation of the states |n〉 introduced in Eq. (18),

and the tensor structure given to the them and to theoperator K in Eq. (20) is of central importance in thederivation of MMFT and is the main focus of our criti-cism.

A linear Diophantine equation of the form in Eq.(19) with unknowns p and ni can always be solvedindependently of the specific integer values n and Ni[51]. In fact, it has infinitely many solutions. For in-stance, given a solution np it is always possible toobtain another solution by redefining the vector n asn′ = (n1 + zN2, n2 − zN1) with an arbitrary integer z.For this reason, it is not possible to uniquely associate asingle vector |np〉 with each vector |n〉 without a specificprescription of which solution to choose. Such prescrip-tion, however, is not given in [27] and is not compatiblewith the tensor structure provided [27].

Problems arising from the ambiguity in the identifica-tion of the vector |np〉 in Eq. (18) become apparent whenconsidering e.g. the scalar product of two states |np〉 and|n′p′〉 that correspond to two solutions np and n′p′.The scalar product 〈np|n′p′〉 vanishes if the two solu-tions are different. However, with the original notationboth states are associated with the same state |n〉 andthe corresponding scalar product 〈n|n〉 does not vanish,which leads to an inconsistency. This problem becomesespecially relevant when considering the expression of theoperator K in Eq. (20), which contains infinitely many

different matrix elements 〈np|K|mq〉 that correspond to

the same matrix element 〈n|K|m〉 of the operator K in

Eq. (11). That is, the operator Kbd in Eq. (20) is in fact

not a mere rewritten version of K in Eq. (11) but rathera different operator.

A central step in the derivation of MMFT is the exis-tence of a unitary transformation relating the operatorsKbd defined in Eq. (20) and

Kd = KB ⊗ 1 + 1⊗ ωp (22)

with

KB = HF ⊗ 1 + 1⊗ ω · n (23)

The existence would follow from a bi-jective relation be-tween |n〉 and |np〉; but since such a relation does notexist, the unitary equivalence between the two operatorsdoes not necessarily hold true.

The notation introduced in Eq. (18) is also employed

to express the time-evolution operator U(t) in Eq. (12)as

∞∑n1,n2,p=−∞

〈n1n2p|e−iKt|000〉einωt. (24)

This expression, however, contains infinitely many dupli-cate terms since there are infinitely many vectors 〈n1n2p|

5

that correspond to the same vector 〈n|, according to theprescription given by the Diophantine equation in Eq.(19). Eq. (24) is thus not a reformulation of the expres-sion of the time-evolution operator in Eq. (12).

The derivation of MMFT [27] achieves the desiredstructure of the time-evolution operator by combiningthe expression for the time-evolution operator U(t) in Eq.

(24) with the expression for the operator K in Eq. (20).

Given the doubts on the unitary equivalence between Kbdand Kd and the correctness of Eq. (24), it seems to usthat the derivation of MMFT is not complete.

Besides a Floquet-like decomposition for quasi-periodicsystems, MMFT also describes a method to calculatethe time-evolution operator by diagonalizing a time-independent operator perturbatively or numerically, ina similar way as described in Eq. (13) for periodic sys-tems [47]. Specifically, it is argued [27] that finding the

unitary transformation that relates Kbd and Kd is essen-tially equivalent to transforming K in Eq. (21) to KB inEq. (23). This method has then been applied in a varietyof fields, leading to successful results [28, 38–40, 42–44].

In the next section we will give an explanation why,despite arguing that the proof of MMFT is not entirelyrigorous and possibly incomplete, this method can stilllead to valid results. We shall do this without imposingany intermediate periodicity in the system but rather bydirectly defining an extended Hilbert space, in an analo-gous way as described in Sec. II A for periodic systems.

IV. QUASI-PERIODIC REDUCIBILITY INFOURIER SPACE

The possibility to express the time-evolution operatorU(t) of quasi-periodically driven systems in a generalizedFloquet decomposition can be formulated in terms of thereducibility of the Schrodinger equation, as described inSec. II A for periodic systems. In the quasi-periodic case,we seek a quasi-periodic unitary UQ(t) that transformsthe operator K(t) = H(t)− i∂t to

UQ(t)K(t)U†Q(t) = HQ − i∂t, (25)

whereHQ = UQ(t)H(t)U†Q(t)−iUP (t)(∂tU†Q(t)) is a time-

independent Hermitian operator and UQ(0) = 1 [52].In Sec. II A, we have described how the transforma-

tion in Eq. (4)—which is known to exist due to Floquet’stheorem—can be solved within a time-independent for-malism using Fourier series. Here, we expand this formal-ism to include quasi-periodic systems and show how thetransformation in Eq. (25) can be similarly formulated interms of the block-diagonalization of a time-independentoperator. With this, we will not aim at proving the exis-tence of the decomposition of the time-evolution operatorin Eq. (2), but rather assume its existence and constructthe corresponding effective Hamiltonian.

Similarly as described in Sec. II for periodic sys-tems, the Fourier coefficients of quasi-periodic states

can be defined as the Fourier components of vectors inH ⊗ L2(Td), where H is the original system’s Hilbertspace and L2(Td) is the space of square-integrable func-tions on a d-dimensional torus. Thereafter, the iso-morphism between the space L2(Td) and the sequencespace l2(Zd) can be employed to work within a time-independent or Fourier formalism. In this manner, quasi-periodic operators B(t) =

∑nBne

in·ωt can be mappedto

B =∑n

Bn ⊗ σn, (26)

were the ladder operators σn =∑

m |m + n〉〈m| are de-fined in terms of a basis |n〉 of the sequence space l2(Zd)and satisfy σn|m〉 = |n + m〉. Similarly, the derivativeoperator −i∂t can be mapped to

D = 1⊗ n · ω, (27)

with the number operator n =∑

n n|n〉〈n| satisfyingn|n〉 = n|n〉 and the commutation relation [n, σm] =mσm. The operator K(t) = H(t) − i∂t can then be as-sociated to the operator

K =∑n

Hn ⊗ σn + 1⊗ n · ω, (28)

which coincides with the operator already introduced inEq. (21). In this way, the transformation in Eq. (25) isthen given by

KB = UQKU†Q = HQ ⊗ 1 + 1⊗ n · ω, (29)

where the transformation UQ has the form

UQ =∑n

Un ⊗ σn, (30)

since it describes a quasi-periodic unitary.The block-diagonalization described in Eq. (29) offers

an alternative formulation of the transformation in Eq.(25), and indeed coincides with the transformation relat-

ing Kbd and Kd defined in Eqs. (20) and (22). That is,even if the general premise of MMFT is not satisfied, itsexplicit application is still correct as long as reducibilityholds.

V. GENERALIZED FLOQUET-MAGNUSEXPANSION

General results of reducibility for first-order differentialequations with quasi-periodic coefficients are not to beexpected [32, 33], but the situation is better understoodif the driving amplitude is small as compared to the norm|ω| of the frequency vector. Using unitless variables τ =|ω|t, which are common in the mathematical literature,the corresponding differential equation reads

i∂τU(τ) =H(τ)

|ω|U(τ) . (31)

6

The regime of the new rescaled Hamiltonian, which isquasi-periodic with frequencies ω/|ω|, is referred to aclose-to-constant, whereas the term fast driving is morecommon in the physics literature. In this regime andunder suitable hypothesis of regularity, non-degeneracyand strong nonresonance of the frequencies, reducible andnon-reducible systems are mixed like Diophantine and Li-ouvillean numbers: most systems are reducible but non-reducible ones are dense [34, 35, 53, 54]. Morover, thegeneralized Floquet decomposition of the time-evolutionoperator in Eq. (2) can be found with any given accuracy,for |ω|−1 sufficiently small, provided it exists [55–57]. Inpractice, this is often done through an expansion in termsof powers of |ω|−1.

In this section, we will derive a generalization of theFloquet-Magnus expansion [45, 58, 59] to quasi-periodicsystems and provide a perturbative exponential expan-sion of the time-evolution operator with the desired Flo-quet representation. This will allow us to identify HQ asthe effective Hamiltonian that captures the dynamics ofthe system in a suitable fast-driving regime.

We start the derivation by reproducing the steps of theregular Floquet-Magnus expansion [45] and introducingthe desired decomposition of the time-evolution operator

U(t) = U†Q(t)e−iHQt (32)

into the Schrodinger equation i∂tU(t) = H(t)U(t), whichyields the differential equation

i∂tU†Q(t) = H(t)U†Q(t)− U†Q(t)HQ. (33)

Then, we define the quasi-periodic Hermitian operatorQ(t) as the generator of the quasi-periodic unitary UQ(t)via the relation

UQ(t) = eiQ(t). (34)

Introducing the expression in Eq. (34) into Eq. (33)and using a power series expansion for the differential ofthe exponential [45, 59, 60], one obtains the non-lineardifferential equation [45]

∂tQ(t) =

∞∑k=0

Bkk!

(−i)kadkQ(t)

(H(t) + (−1)k+1HQ

),(35)

where Bk denote the Bernoulli numbers and ad is theadjoint action defined via adkAB = [A, adk−1

A B] for k ≥ 1

and ad0AB = B.

The next step in the derivation is to consider a seriesexpansion for the operators HQ and Q(t) of the form

HQ =

∞∑n=1

H(n)Q (36)

Q(t) =

∞∑n=1

Q(n)(t), (37)

with Q(n)(0) = 0 and where the superscript indicates theorder of the expansion. After introducing the series in

Eqs. (36) and (37) into Eq. (35) and equating the termswith the same order, one obtains the differential equation

∂tQ(n)(t) = A(n)(t)−H(n)

Q , (38)

with A(1)(t) = H(t) and

A(n)(t) =

n−1∑k=1

Bkk!

(X(n)k (t) + (−1)k+1Y

(n)k ) (39)

for n ≥ 2. The operators X(n)k (t) and Y

(n)k (t) in Eq. (39)

are given recursively by

X(n)k (t) =

n−k∑m=1

[Q(m)(t), X(n−m)k−1 (t)] (40)

Y(n)k (t) =

n−k∑m=1

[Q(m)(t), Y(n−m)k−1 (t)] (41)

for 1 ≤ k ≤ n − 1, with X(1)0 = −iH(t), X

(n)0 = 0 for

n ≥ 2, and Y(n)0 = −iH(n)

Q for all n.An important feature of the differential equation in

Eq. (38) is the structure of the operator A(n)(t), whichonly contains terms involving the Hamiltonian H(t) or

operators Q(m)(t) and H(m)Q of a lower order, i.e. with

m < n. This allows to solve Eq. (38) by just integratingthe right hand side of the equation, which leads to

Q(n)(t) =

∫ t

0

(A(n)(t)−H(n)

Q

)dt. (42)

Moreover, even though Eq. (38) describes a differentialequation for Q(n)(t), the solutions for both Q(n)(t) and

H(n)Q can be inferred from it by imposing suitable condi-

tions on the time dependence of Q(n)(t). In the periodic

case, for example, the operators H(n)Q are fixed by the

requirement that Q(n)(t) is a periodic operator [45].

In the quasi-periodic case, we can determine H(n)Q by

exploiting the quasi-periodicity of Q(n)(t) and A(n)(t).This essentially results from the fact that, in orderfor the integral of a quasi-periodic operator O(t) =∑

nOnein·ωt to be quasi-periodic, it must satisfy that

O0 = limT→∞1T

∫ T0O(t)dt = 0. As a consequence, in

order for Q(n)(t) in Eq. (42) to be quasi-periodic, H(n)Q

must read

H(n)Q = lim

T→∞

1

T

∫ T

0

A(n)(t)dt. (43)

Eqs. (42) and (43) can be solved for any n > 1 providedthat the solutions for m < n are known. Since they canbe readily solved for n = 1, Eqs. (42) and (43) thuscontain the necessary information to recursively deriveall the terms in the expansions of Q(t) and HQ in Eqs.(36) and (37), respectively.

7

After performing the integrations in Eqs. (42) and(43), the first two terms terms of the series for HQ be-come

H(1)Q = H0 (44)

H(2)Q =

1

2

∑n6=0

[Hn, H−n]

ω · n+∑n6=0

[H0, Hn]

ω · n, (45)

whereHn are the Fourier coefficients of the quasi-periodicHamiltonian, as defined in Eq. (1). Similarly, the firsttwo terms of Q(t) read

Q(1)(t) = −i∑n6=0

Hn

n · ω(ein·ωt − 1) (46)

Q(2)(t) =i

2

∑n6=0

[H0, Hn]

(n · ω)2(ein·ωt − 1)

+i

2

∑n6=0;m 6=−n

[Hn, Hm]

n · ω (n + m) · ω(ei(n+m)·ωt − 1)

+i

2

∑n6=0;m 6=0

[Hn, Hm]

n · ω m · ω(eim·ωt − 1) . (47)

Consistently with the periodic case, the results obtainedhere reduce to the regular Floquet-Magnus expansionwhen the frequency vector ω contains only one element.Moreover, by using the Baker-Campbell-Hausdorff for-mula [61], one can verify that the expressions in Eqs.(44)-(47) coincide with the first terms of the regularMagnus expansion [59], which applies for general time-dependent systems. This formal expansion can also belinked [62] to the method of averaging for quasi-periodicsystems [55] to obtain exponentially small error estimatesin the quasi-periodic case.

The expansion of the operators HQ and Q(t) intro-duced in Eqs. (36) and (37) can be interpreted as a se-ries expansion in powers of |ω|−1 such that, in a suitablefast-driving regime, the lowest-order terms of the seriesare the most relevant to describe the dynamics of the sys-tem. Even though the convergence of the quasi-periodicFloquet-Magnus expansion is in general not guaranteedand requires further investigations, this permits us toidentify effective Hamiltonian analogously as done for pe-riodic systmes.

In fast-driving regimes where the fundamental drivingfrequencies are the largest energy scales of the system,the two unitaries UQ(t) and e−iHQt of the time-evolutionoperator in Eq. (32) capture two disctinct behaviours ofthe system’s dynamics. On the one hand, the unitaryUQ(t) describes fast quasi-periodic fluctuations dictatedby the fast frequencies ω. On the other hand, the oper-ator e−iHQt captures the slower dynamics of the systemcharacterized by the internal energy scales of HQ, whichcan be thus identified as the effective Hamiltonian of thesystem.

VI. QUASI-PERIODICALLY DRIVEN LAMBDASYSTEM

In this section, we will illustrate with a quasi-periodically driven Lambda system the possibility to ap-proximate the dynamics of quasi-periodically driven sys-tems in terms of a truncation of the effective HamiltonianHQ.

The Lambda system describes an atomic three energy-level system with two ground states |1〉 and |2〉 that arecoupled to an excited state |3〉 via a time-dependent laserfield. The time-dependent coupling allows one to indi-rectly mediate a transition between the states |1〉 and |2〉without significantly populating the excited state and,in this way, overcome the impossibility to drive a di-rect transition between the two degenerate ground states.This method also permits the implementation of non-trivial phases in the tunneling rate of particles [63, 64]and constitutes a building block in many quantum sim-ulations [13, 16, 20, 65].

The Hamiltonian of the Lambda system in an interac-tion picture reads

H(t) = f(t)|3〉 (〈1|+ 〈2|) + H.c, (48)

where f(t) is usually a periodic function but, here, weconsider it to be quasi-periodic; i.e. of the form

f(t) =∑n

fnein·ωt, (49)

with a frequency vector ω and Fourier components fn.Moreover, we require the static Fourier component tovanish, i.e. f0 = 0, in order to ensure that the dominantdynamics of the system do not yield transitions betweenthe ground states and the excited state.

With the Hamiltonian of the quasi-periodically drivenLambda system in Eq. (48), the first two terms of theeffective Hamiltonian expansion in Eqs. (44) and (45)

become H(1)Q = 0 and

H(2)Q = Ωeff ((|1〉+ |2〉)(〈1|+ 〈2|)− 2|3〉〈3|) , (50)

respectively. As the first term vanishes, the leading order

term of the effective Hamiltonian is thus given by H(2)Q ,

which describes transitions between the ground states ofthe system with a rate

Ωeff =∑n

|fn|2

n · ω. (51)

In order to illustrate the possibility to approximate thedynamics of the system in terms of a truncation of HQ,we compare in Figs. 1 and 2 matrix elements of a numer-ically exact calculation of the time-evolution operator ofthe system U(t) and the effective time-evolution operator

Ueff(t) = e−iH(2)Q t (52)

for different driving functions f(t). Specifically, we dis-play the transition probabilities P12(t) = |〈1|U(t)|2〉|2

8

(a)

Tran

sitio

n pr

obab

ility

P12

P e↵12

0 100 200 300 4000.0

0.2

0.4

0.6

0.8

1.0

Tran

sitio

n pr

obab

ility

P12

P e↵12

!1t0 100 200 300 4000.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0

!1t0 100 200 300 400

0 100 200 300 400

(b)

Tran

sitio

n pr

obab

ility

P12

P e↵12

0 100 200 300 4000.0

0.2

0.4

0.6

0.8

1.0

Tran

sitio

n pr

obab

ility

P12

P e↵12

!1t0 100 200 300 4000.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0

!1t0 100 200 300 400

0 100 200 300 400

FIG. 1. Comparison between the exact and effective transition probabilities P12(t) = |〈1|U(t)|2〉|2 and P eff12 (t) = |〈1|Ueff(t)|2〉|2

for the periodically (a) and quasi-periodically (b) driven Lambda system. In (a), a periodic driving function f(t) = Ωeiω1t with

Ω/ω1 = 0.1(1 +√

2/2)1/2 is considered. In (b), the results correspond to a quasi-periodic function f(t) = Ω(eiω1t + ei√

2ω1t)with Ω/ω1 = 0.1. The parameters of the driving function in (a) and (b) are such that they lead to the same effective rate Ωeff

in Eq. (51).

and P eff12 (t) = |〈1|Ueff(t)|2〉|2, which describe the exact

and effective transitions between the ground states of theLambda system.

In Fig. 1, we compare the performance of the drivenLambda system for a periodic and quasi-periodic drivingfunctions in a moderately fast driving regime. In Fig. 1(a), a periodic driving is considered, which yields exactdynamics that exhibit fast regular fluctuations aroundthe slower effective dynamics. On the contrary, we showin Fig. 1 (b) how a quasi-periodic driving leads to apattern with seemingly erratic fluctuation around the ef-fective dynamics. In the regime where the fluctuationscan be neglected, however, their regularity is irrelevant.This supports the view that, since quasi-periodic func-tions provide a more general parametrization of the driv-ing protocol, quasi-periodically driven quantum systemshave the potential to expand the accessible effective dy-namics in a variety of experimental setups [66].

Another aspect that is apparent from Fig. 1 is thedrift between the exact and effective dynamics. This isnot a characteristic feature of quasi-periodically drivensystems but rather results from the truncation of the op-erator HQ. Including higher order terms in the expansionof HQ or considering a faster driving regime, the approx-imation would be improved and the exact and effectivedynamics of the system would better overlap for longertimes. Indeed, in Fig. 2 we consider a quasi-periodicfunction with higher frequencies and observe that theeffective time-evolution operator in Eq. (52) approxi-mates better the exact dynamics of the system for longertimes. This highlights the possibility to use the gener-alized Floquet-Magnus expansion derived in Sec. V inorder to derive time-independent effective Hamiltoniansthat capture well the dynamics of quasi-periodic systemsin a suitable fast-driving regime.

0 500 1000 1500 2000 2500 3000 3500 40000.0

0.2

0.4

0.6

0.8

1.0

Tran

sitio

n pr

obab

ility

P12

P e↵12

!1t0.0

0.2

0.4

0.6

0.8

1.0

0 1000 2000 3000 4000

FIG. 2. Plot of the transition probabilities P12(t) =|〈1|U(t)|2〉|2 and P eff

12 (t) = |〈1|Ueff(t)|2〉|2 as a function of timefor a quasi-periodic Lambda system with f(t) = Ω(eiω1t +

ei√

2ω1t) and Ω/ω1 = 0.05.

VII. CONCLUSIONS

Despite concerted efforts towards generalisation of Flo-quet’s theorem for quasi-periodic systems, there are stillmany open questions regarding the existence of Floquet-like decompositions. Although a rigorous footing is notcomplete, effective Hamiltonians can be constructed forweakly and/or rapidly driven quantum systems.

Since the restriction to periodic driving naturally im-poses restrictions on the effective Hamiltonians that canbe achieved, the use of quasi-periodic Hamiltonians isa promising route for quantum simulations. The in-creased freedom in accessible time-dependencies makesquasi-periodic driving a highly interesting basis for theidentification of accurate implementations of effective dy-namics by means of optimal control. As such, one can ex-pect that quasi-periodic driving will find increased appli-cation in quantum simulations, and that the increased in-terest in quantum physics will trigger activities in math-ematics towards the existence of Floquet-like decompo-sitions and the convergence of perturbative expansions.

9

ACKNOWLEDGMENTS

A.V. and F.M. acknowledge financial support by theEuropean Research Council within the project ODYC-QUENT. The research of J.P. is supported by grants2014-SGR-00504 and MTM2015-65715-P.

Appendix: The propagator in Floquet theory

Here, we show that

U(t) =∑n

〈n|e−iKt|0〉einωt, (A.1)

as given in Eq.(12) is indeed the propagator induced byH(t), i.e. that it satisfies the Schrodinger equation with

the initial condition U(0) = 1.The time-derivative of Eq.(A.1) reads

i∂tU(t) =∑n

〈n|(K − nω)e−iKt|0〉einωt . (A.2)

Using the explicit form

K =∑n

Hn ⊗ σn + 1⊗ ωn. (A.3)

and

〈n|n = 〈n|n , (A.4)

Eq. (A.2) reduces to

i∂tU(t) =∑nm

〈n|Hm ⊗ σme−iKt|0〉einωt (A.5)

=∑nm

Hm〈n|σme−iKt|0〉einωt (A.6)

=∑nm

Hm〈n−m|e−iKt|0〉einωt . (A.7)

Replacing the summation index n by n+m leads to

i∂tU(t) =∑nm

Hm〈n|e−iKt|0〉ei(n+m)ωt (A.8)

=∑m

Hmeimωt

∑n

〈n|e−iKt|0〉einωt (A.9)

= H(t)U(t) . (A.10)

The initial condition U(0) = 1 results directly from

〈n|e−iK0|0〉 = 1.

[1] M. Shapiro and P. Brumer, Int. Rev. Phys. Chem., 13,187 (1994).

[2] W. Becker, X. Liu, P. J. Ho, and J. H. Eberly, Rev. Mod.Phys., 84, 1011 (2012).

[3] G. Engel, T. Calhoun, E. Read, T.-K. Ahn, T. Mancal,Y.-C. Cheng, R. Blankenship, and G. Fleming, Nature,446, 782 (2007).

[4] H. Ichikawa, S. Nozawa, T. Sato, A. Tomita,K. Ichiyanagi, M. Chollet, L. Guerin, N. Dean, A. Cav-alleri, S.-i. Adachi, T.-h. Arima, H. Sawa, Y. Ogimoto,M. Nakamura, R. Tamaki, K. Miyano, and S.-y. Koshi-hara, Nat Mater, 10, 101 (2011).

[5] Y. Tokura, J. Phys. Soc. Jap., 75, 011001 (2006).[6] M. Bukov, L. D’Alessio, and A. Polkovnikov, Adv. Phys.,

64, 139 (2015).[7] N. Goldman and J. Dalibard, Phys. Rev. X, 4, 031027

(2014).[8] L. H. Eliasson, Comm. Math. Phys., 146, 447 (1992).

[9] G. Floquet, Ann. Ecole Norm. Sup., 12, 47 (1883).[10] B. J. Keay, S. Zeuner, S. J. Allen, K. D. Maranowski,

A. C. Gossard, U. Bhattacharya, and M. J. W. Rodwell,Phys. Rev. Lett., 75, 4102 (1995).

[11] K. W. Madison, M. C. Fischer, R. B. Diener, Q. Niu,and M. G. Raizen, Phys. Rev. Lett., 81, 5093 (1998).

[12] H. Lignier, C. Sias, D. Ciampini, Y. Singh, A. Zenesini,O. Morsch, and E. Arimondo, Phys. Rev. Lett., 99,220403 (2007).

[13] Y.-J. Lin, K. Jimenez-Garcia, and I. B. Spielman, Na-ture, 471, 83 (2011).

[14] G. Jotzu, M. Messer, F. Gorg, D. Greif, R. Desbuquois,and T. Esslinger, Phys. Rev. Lett., 115, 073002 (2015).

[15] J. Struck, C. Olschlager, M. Weinberg, P. Hauke, J. Si-

monet, A. Eckardt, M. Lewenstein, K. Sengstock, andP. Windpassinger, Phys. Rev. Lett., 108, 225304 (2012).

[16] M. Aidelsburger, M. Atala, M. Lohse, J. T. Barreiro,B. Paredes, and I. Bloch, Phys. Rev. Lett., 111, 185301(2013).

[17] H. Miyake, G. A. Siviloglou, C. J. Kennedy, W. C. Bur-ton, and W. Ketterle, Phys. Rev. Lett., 111, 185302(2013).

[18] C. V. Parker, L.-C. Ha, and C. Chin, Nat. Phys., 9, 769(2013).

[19] G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat, T.Uehlinger, D. Greif, and T. Esslinger, Nature, 515, 237(2014), ISSN 0028-0836.

[20] Aidelsburger, M., Lohse, M., Schweizer, C., Atala, M.,Barreiro, J. T., Nascimbene, S., Cooper, N. R., Bloch,I., and Goldman, N., Nat. Phys., 11, 162 (2015), ISSN1745-2473.

[21] A. Verdeny, L. Rudnicki, C. A. Muller, and F. Mintert,Phys. Rev. Lett., 113, 010501 (2014).

[22] P. Hauke, O. Tieleman, A. Celi, C. Olschlager, J. Si-monet, J. Struck, M. Weinberg, P. Windpassinger,K. Sengstock, M. Lewenstein, and A. Eckardt, Phys.Rev. Lett., 109, 145301 (2012).

[23] A. Verdeny and F. Mintert, Phys. Rev. A, 92, 063615(2015).

[24] E. Neumann and A. Pikovsky, EPJD, 26, 219 (2002).[25] R. Gommers, S. Denisov, and F. Renzoni, Phys. Rev.

Lett., 96, 240604 (2006).[26] D. Cubero and F. Renzoni, Phys. Rev. E, 86, 056201

(2012).[27] T.-S. Ho, S.-I. Chu, and J. V. Tietz, Chem. Phys. Lett.,

96, 464 (1983), ISSN 0009-2614.

10

[28] S.-I. Chu and D. A. Telnov, Phy. Rep., 390, 1 (2004),ISSN 0370-1573.

[29] A. Jorba and C. Simo, Journal of Differential Equations,98, 111 (1992).

[30] R. Krikorian, Ann. of Math., 154, 269 (2001).[31] A. Avila and R. Krikorian, Annals of Mathematics, 164,

911 (2006).[32] L. H. Eliasson, in Doc. Math., Extra Vol. II (1998) pp.

779–787 (electronic).[33] L. H. Eliasson, in XIVth International Congress on Math-

ematical Physics (World Scientific, 2006) pp. 195–205.[34] N. Karaliolios, To appear in Memoires SMF, 138

(2015).[35] L. H. Eliasson, Ergodic Theory Dynam. Systems, 22,

1429 (2002).[36] R. Krikorian, Ergodic Theory Dynam. Systems, 19, 61

(1999).[37] H.-L. Her and J. You, Journal of Dynamics and Differ-

ential Equations, 20, 831 (2008).[38] T.-S. Ho and S.-I. Chu, J. Phys. B, 17, 2101 (1984).[39] T.-S. Ho and S.-I. Chu, Phys. Rev. A, 31, 659 (1985).[40] T.-S. Ho and S.-I. Chu, Phys. Rev. A, 32, 377 (1985).[41] M. Leskes, R. S. Thakur, P. K. Madhu, N. D. Kurur,

and S. Vega, J. Chem. Phys., 127, 024501 (2007), doi:http://dx.doi.org/10.1063/1.2746039.

[42] S.-K. Son and S.-I. Chu, Phys. Rev. A, 77, 063406 (2008).[43] D. Zhao, C.-W. Jiang, and F.-l. Li, Phys. Rev. A, 92,

043413 (2015).[44] A. M. Forney, S. R. Jackson, and F. W. Strauch, Phys.

Rev. A, 81, 012306 (2010).[45] F. Casas, J. A. Oteo, and J. Ros, J. Phys. A, 34, 3379

(2001).[46] The periodic unitary UP (t) and the Hamiltonian HF are,

however, not uniquely defined. In particular, the eigen-values of HF are only defined up to multiples of ω.

[47] J. H. Shirley, Phys. Rev., 138, B979 (1965).[48] H. Sambe, Phys. Rev. A, 7, 2203 (1973).[49] A. Verdeny, A. Mielke, and F. Mintert, Phys. Rev. Lett.,

111, 175301 (2013).[50] For convenience, we use the notation introduced in Sec.

II, which differs from the one used in [27]. In particu-

lar, the operator K is denoted by HF in [27] and it is

represented in terms of a matrix.[51] A Diophantine equation n =

∑di=1 niNi with unknowns

ni can be solved if and only if the greatest common di-visor gcd(N1, · · · , Nd) divides n [? ]. Thus, by appropri-ately choosing the variable p, the Diophantine equationin Eq. (19) can always be solved.

[52] Similarly to the operator HF introduced in Eq. (4), theeigenvalues of HQ are only defined up n ·ω, with a vectorof integers n and the frequency vector ω of the Hamilto-nian H(t).

[53] R. Krikorian, Annals of Mathematics, 154, 269 (2001),ISSN 0003486X.

[54] K. Fraczek, 139 (2004).[55] C. Simo, “Hamiltonian mechanics: Integrability and

chaotic behavior,” (Springer US, Boston, MA, 1994)Chap. Averaging under Fast Quasiperiodic Forcing, pp.13–34, ISBN 978-1-4899-0964-0.

[56] D. V. Treshchev, Translations of the American Mathe-matical Society-Series 2, 168, 91 (1995).

[57] A. Jorba, R. Ramırez-Ros, and J. Villanueva, SIAMJournal on Mathematical Analysis, 28, 178 (1997).

[58] S. Blanes, F. Casas, J. Oteo, and J. Ros, Phys. Rep.,470, 151 (2009), ISSN 0370-1573.

[59] W. Magnus, Comm. P. and App. Math., 7, 649 (1954),ISSN 1097-0312.

[60] R. M. Wilcox, Journal of Mathematical Physics, 8, 962(1967).

[61] V. Varadarajan, Lie Groups, Lie Algebras, and TheirRepresentations (Springer-Verlag New York, 1984).

[62] P. Chartier, A. Murua, and J. Sanz-Serna, Discrete andContinuous Dynamical Systems, 32, 1 (2012).

[63] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah,W. S. Bakr, and M. W. Zwierlein, Phys. Rev. Lett., 109,095302 (2012).

[64] P. Wang, Z.-Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai,H. Zhai, and J. Zhang, Phys. Rev. Lett., 109, 095301(2012).

[65] Y.-J. Lin, R. L. Compton, K. Jimenez-Garcia, J. V.Porto, and I. B. Spielman, Nature, 462, 628 (2009).

[66] C. Yuce, Europhys. Lett., 103, 30011 (2013).