7
Optimal Control of Effective Hamiltonians Albert Verdeny, 1,2 Lukasz Rudnicki, 1,3 Cord A. Müller, 4,5 and Florian Mintert 1,2 1 Freiburg Institute for Advanced Studies, Albert-Ludwigs-Universität, Albertstrasse 19, 79104 Freiburg, Germany 2 Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom 3 Center for Theoretical Physics, Polish Academy of Sciences, Aleja Lotników 32/46, 02-668 Warsaw, Poland 4 Centre for Quantum Technologies, National University of Singapore, Singapore 117543, Singapore 5 Department of Physics, University of Konstanz, 78457 Konstanz, Germany (Received 6 February 2014; published 2 July 2014) We present a systematic scheme for the optimization of quantum simulations. Specifically, we show how polychromatic driving can be used to significantly improve the driving of Raman transitions in the Lambda system, which opens new possibilities for controlled driving-induced effective dynamics. DOI: 10.1103/PhysRevLett.113.010501 PACS numbers: 03.67.Ac, 02.30.Yy, 37.10.Jk In the past few years, one of the most active and promising research fields has been the design of quantum simulators, i.e., engineered controllable quantum systems utilized to mimic the dynamics of other systems. With this, new insight is expected to be gained in a variety of phenomena like high-temperature fractional quantum Hall states [1], (non-)Abelian gauge fields [2,3], and even relativistic effects [4,5]. Driven systems provide a powerful tool to simulate desired effective dynamics. An important example is laser-assisted Raman transitions between different elec- tronic states and/or localized states of trapped atoms, which is a central pillar in a large number of quantum simulations. Because the direct coupling between low-lying energy states via dipole transitions is often forbidden by selection rules, an intermediate auxiliary state with higher energy is usually used to mediate the coupling. This so-called Lambda system is then specifically configured to imprint phases required to realize various spin-orbit couplings [6,7] or to simulate the effect of gauge fields [8,9]. Other prominent examples of driving-induced effective dynamics include shaken lattices [1012], lattices with modulated interactions [13], or driven graphene [14]. Even though driven systems provide a powerful approach to perform quantum simulations, they often rely on approximations that currently situate them still far from the ideal quantum simulator. For instance, in Raman transitions via a three-level Lambda system, the driving pulse produces an undesired population of the excited state. These deviations between the desired and simulated dynamics accumulate during the evolution and become considerable after a sufficiently long time. From the experimental side, however, spectacular progress has been made in the manipulation and control of quantum systems [15,16], so that accurate theoretical tools to choose the proper driving are necessary. The field of optimal control theory [17,18] aims at such precise manipulation but, so far, it has primarily targeted properties at single instances in time [1922] whereas we are rather concerned with the behavior of a system during a continuous time window. In this Letter, we provide a general systematic approach to improve quantum simulations by using pulse shaping techniques of optimal control theory. We discuss in detail the optimal control of the Lambda system and rigorously show how an appropriately chosen polychromatic driving can significantly improve Raman transitions. As a result, we do not only provide a proof of principle for the optimal control of effective Hamiltonians but also optimize a building block used in a large variety of quantum simulations. Consider the target dynamics U tg ¼ e iH tg t generated by the target Hamiltonian H tg that we wish to simulate using a time-periodic driving Hamiltonian HðtÞ¼ Hðt þ T Þ. Its time-evolution operator UðtÞ¼ T exp½i R t 0 Hðt 0 Þdt 0 then admits the Floquet decomposition [23] UðtÞ¼ ~ UðtÞU eff ðtÞ: ð1Þ Here ~ UðtÞ is a T -periodic unitary satisfying ~ Uð0Þ¼ 1, U eff ðtÞ¼ e iH eff t , and H eff is a time-independent effective Hamiltonian defined via UðT Þ¼ e iH eff T . ~ U describes fluctuations around the envelope evolution U eff . These fluctuations become negligible if the dominant energy scale Ω d of HðtÞ is sufficiently small compared to the driving frequency ω ¼ 2π=T. In this case the low-energy or long- time dynamics of the periodically driven system is well described by H eff , which in turn should be chosen to match the target Hamiltonian H tg to be simulated. Suppose now that the driving HðtÞ contains a set ff n g of control parameters. Our aim is to tune these parameters such that the dynamics U resembles the target dynamics U tg as well as possible. Different choices of ff n g can result in similar effective dynamics, but produce different fluctuations. In order to ensure the optimal simulation of a given target Hamiltonian H tg with the least fluctuations, our scheme therefore consists of the following. (i) Identifying the dependence of the effective Hamiltonian H eff on the control PRL 113, 010501 (2014) PHYSICAL REVIEW LETTERS week ending 4 JULY 2014 0031-9007=14=113(1)=010501(5) 010501-1 © 2014 American Physical Society

Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

Optimal Control of Effective Hamiltonians

Albert Verdeny,1,2 Łukasz Rudnicki,1,3 Cord A. Müller,4,5 and Florian Mintert1,21Freiburg Institute for Advanced Studies, Albert-Ludwigs-Universität, Albertstrasse 19, 79104 Freiburg, Germany

2Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom

3Center for Theoretical Physics, Polish Academy of Sciences, Aleja Lotników 32/46, 02-668 Warsaw, Poland

4Centre for Quantum Technologies, National University of Singapore, Singapore 117543, Singapore

5Department of Physics, University of Konstanz, 78457 Konstanz, Germany

(Received 6 February 2014; published 2 July 2014)

We present a systematic scheme for the optimization of quantum simulations. Specifically, we show howpolychromatic driving can be used to significantly improve the driving of Raman transitions in the Lambdasystem, which opens new possibilities for controlled driving-induced effective dynamics.

DOI: 10.1103/PhysRevLett.113.010501 PACS numbers: 03.67.Ac, 02.30.Yy, 37.10.Jk

In the past few years, one of the most active andpromising research fields has been the design of quantumsimulators, i.e., engineered controllable quantum systemsutilized to mimic the dynamics of other systems. Withthis, new insight is expected to be gained in a variety ofphenomena like high-temperature fractional quantum Hallstates [1], (non-)Abelian gauge fields [2,3], and evenrelativistic effects [4,5].Driven systems provide a powerful tool to simulate

desired effective dynamics. An important example islaser-assisted Raman transitions between different elec-tronic states and/or localized states of trapped atoms, whichis a central pillar in a large number of quantum simulations.Because the direct coupling between low-lying energystates via dipole transitions is often forbidden by selectionrules, an intermediate auxiliary state with higher energyis usually used to mediate the coupling. This so-calledLambda system is then specifically configured to imprintphases required to realize various spin-orbit couplings[6,7] or to simulate the effect of gauge fields [8,9]. Otherprominent examples of driving-induced effective dynamicsinclude shaken lattices [10–12], lattices with modulatedinteractions [13], or driven graphene [14].Even though driven systems provide a powerful

approach to perform quantum simulations, they often relyon approximations that currently situate them still far fromthe ideal quantum simulator. For instance, in Ramantransitions via a three-level Lambda system, the drivingpulse produces an undesired population of the excited state.These deviations between the desired and simulateddynamics accumulate during the evolution and becomeconsiderable after a sufficiently long time. From theexperimental side, however, spectacular progress has beenmade in the manipulation and control of quantum systems[15,16], so that accurate theoretical tools to choose theproper driving are necessary. The field of optimal controltheory [17,18] aims at such precise manipulation but, so far,it has primarily targeted properties at single instances in

time [19–22] whereas we are rather concerned with thebehavior of a system during a continuous time window.In this Letter, we provide a general systematic approach

to improve quantum simulations by using pulse shapingtechniques of optimal control theory. We discuss in detailthe optimal control of the Lambda system and rigorouslyshow how an appropriately chosen polychromatic drivingcan significantly improve Raman transitions. As a result,we do not only provide a proof of principle for the optimalcontrol of effective Hamiltonians but also optimize abuilding block used in a large variety of quantumsimulations.Consider the target dynamics Utg ¼ e−iHtgt generated by

the target Hamiltonian Htg that we wish to simulate using atime-periodic driving Hamiltonian HðtÞ ¼ Hðtþ TÞ. Itstime-evolution operator UðtÞ ¼ T exp½−i

R

t0Hðt0Þdt0 then

admits the Floquet decomposition [23]

UðtÞ ¼ ~UðtÞUeffðtÞ: ð1Þ

Here ~UðtÞ is a T-periodic unitary satisfying ~Uð0Þ ¼ 1,UeffðtÞ ¼ e−iHeff t, and Heff is a time-independent effectiveHamiltonian defined via UðTÞ ¼ e−iHeffT . ~U describesfluctuations around the envelope evolution Ueff . Thesefluctuations become negligible if the dominant energy scaleΩd of HðtÞ is sufficiently small compared to the drivingfrequency ω ¼ 2π=T. In this case the low-energy or long-time dynamics of the periodically driven system is welldescribed by Heff, which in turn should be chosen to matchthe target Hamiltonian Htg to be simulated. Suppose nowthat the driving HðtÞ contains a set ffng of controlparameters. Our aim is to tune these parameters such thatthe dynamics U resembles the target dynamics Utg as wellas possible. Different choices of ffng can result in similareffective dynamics, but produce different fluctuations. Inorder to ensure the optimal simulation of a given targetHamiltonian Htg with the least fluctuations, our schemetherefore consists of the following. (i) Identifying thedependence of the effective HamiltonianHeff on the control

PRL 113, 010501 (2014) P HY S I CA L R EV I EW LE T T ER Sweek ending4 JULY 2014

0031-9007=14=113(1)=010501(5) 010501-1 © 2014 American Physical Society

Page 2: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

parameters ffng. Typically this can only be achievedin an approximate manner, where Heff ¼

P

rk¼0

HðkÞeff þ

OðΩdϵrþ1Þ is known up to order r of the small parameter

ϵ ¼ Ωd=ω ≪ 1. (ii) Constraining ffng so that Htg ¼ Heffto the same order r. (iii) Minimizing the target functional

F ¼1

T

Z

T

0

∥UðtÞ −UtgðtÞ∥2dt; ð2Þ

where jj · jj2 ¼ Trð·†·Þ is the Hilbert-Schmidt norm, underthe constrained control parameters allowed by (ii). Since Fshall also be approximated to the order r consistentwith (i)–(ii) we can evaluate (2) using the substitutionU ¼ ~UUtg. As a result ∥U − Utg∥

2 ¼ Tr½21 − ~U − ~U†, sothat F depends solely on the eigenvalues of ~U [24].In order to calculate Heff , UðtÞ, and F , we use the

Magnus expansion [25,26] and write UðtÞ ¼ e−iMðtÞ

as the exponential of a time-dependent operatorMðtÞ ¼

P

k¼1MkðtÞ. The first two terms of this series

are M1ðtÞ ¼R

t0Hðt1Þdt1 and

M2ðtÞ ¼ −i

2

Z

t

0

dt1

Z

t1

0

dt2½Hðt1Þ; Hðt2Þ: ð3Þ

The kth-order Magnus operator MkðtÞ contains k-fold timeintegrals of k − 1 nested commutators. When HðtÞ is T

periodic, MkðtÞ is exactly of order ϵk. Thus, H

ðkÞeff ¼

Mkþ1ðTÞ=T ∼Ωdϵk, since the factor 1=T ¼ ω=2π reduces

the order of expansion by one. In this manner, Heff , UðtÞ,and consequently F are determined up to order r, and ourscheme (i)–(iii) yields parameters ffng that ensure theoptimal simulation of Htg.In the following, we exemplify the method described

above with a case study of the degenerate Lambda systemwhere j1i and j2i denote the two ground states and j3i theexcited state. The target Hamiltonian

Htg ¼ −Ωtgðj1ih2j þ j2ih1jÞ ð4Þ

generates Raman transitions within the ground-state mani-fold at a rate Ωtg (the overall sign is chosen for laterconvenience). Our aim is to simulate this dynamics bydriving the transitions j1i↔j3i and j2i↔j3iwith a suitablymodulated Rabi frequency. In the interaction picture, wherethe dynamics induced by the static Hamiltonian is absorbedin the state vectors, the driving Hamiltonian takes the form

HðtÞ ¼ fðtÞð1þ e−in0ΔtÞðj1ih3j þ j2ih3jÞ þ H:c: ð5Þ

Importantly, we assume that the driving pulse

fðtÞ ¼X

N

n¼1

fne−inωt ð6Þ

is written as a general Fourier series in terms of ω, thefundamental frequency of driving. Since we do not want the

optimization to rely on strong intensity and fast frequen-cies, the maximal frequency in the above pulse is thedetuning Δ ¼ Nω, i.e., the difference of the resonancefrequency of the Lambda system and the frequency of thedriving carrier. That means that no Fourier componentswith frequency larger than Δ need to be generated. HðtÞdefined in Eq. (5) should be periodic with periodT ¼ 2π=ω, which is the case if Δ is a fraction of twicethe driving carrier frequency, such that n0 is a integer. In therotating wave approximation, one would neglect thecounterrotating contribution e−in0Δt in Eq. (5); we considerthe general case and keep this term.Let us first discuss the simplest example of a mono-

chromatic (MC) driving fðtÞ ¼ f1e−iΔt at constant Rabi

frequency by taking N ¼ 1 in Eq. (6). Since the MagnusoperatorM2lðtÞ of the even order contains even products of

HðtÞ, the corresponding effective Hamiltonian Hð2l−1Þeff has

the desired structure of Htg with matrix elements thatcouple the ground states [27]. Choosing

jf1j2 ¼ ΩtgΔ

1þ n0

2þ n0; ð7Þ

the constraintHð1Þeff ¼ Htg can be fulfilled to first order since

Hð0Þeff ¼ 0. The second-order term H

ð2Þeff , on the other hand,

will generate undesired transitions to the upper level viacubic powers of HðtÞ. With f1 already fixed, one cannot

impose Hð2Þeff ¼ 0, so that one always ends up with an

unwanted population in the excited state—except in theideal limit of very strong, far-detuned driving jf1j, Δ → ∞

at fixedΩtg. Thus, with only one frequency, one can neitheraccurately realize the desired unitary ground-state dynam-ics nor simultaneously minimize the fluctuations.Let us then take advantage of the general pulse, Eq. (6),

and implement the first constraint Hð1Þeff ¼ Htg with

Ωtg ¼1

ω

X

N

n¼1

jfnj2

~nð1Þ; ð8Þ

where ~nðpÞ−1 ≡ n−p þ ðnþ Nn0Þ−p. [In the rotating wave

approximation n0 → ∞ this simplifies to ~nðpÞ ¼ np.]Pushing the Magnus expansion to third order, we can

now require Hð2Þeff ¼ 0 through the second constraint

0 ¼X

N

n¼1

fn~nð1Þ

: ð9Þ

The target functional to be minimized reads [24]

F ð2Þ ¼4

ω2

X

N

n¼1

jfnj2

~nð2Þ: ð10Þ

We can solve the optimization problem now analytically byintroducing two Lagrange multipliers λ1 ∈ R, λ2 ∈ C for

PRL 113, 010501 (2014) P HY S I CA L R EV I EW LE T T ER Sweek ending4 JULY 2014

010501-2

Page 3: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

the two constraints, Eqs. (8) and (9), respectively. Theoptimal pulse parameters are found to be

fn ¼ ωλ2

~nð1Þ

~nð2Þ− λ1

−1

: ð11Þ

The Lagrange multipliers are determined by inserting thissolution into the constraints, Eqs. (8) and (9). DividingEq. (11) by λ2 shows that fn=λ2 is real, and thus all fn aswell as λ2 can be taken as real. Using Eqs. (8) and (9), thetarget functional equation (10) can be rewritten as [24]F ð2Þ ¼ 4λ1Ωtg=ω, such that the global minimum of thefluctuations is found with the minimal root λ1 of Eq. (9)with Eq. (11) inserted. At a given value of ω, one should beable to suppress fluctuations more efficiently using morefrequencies, and indeed, as N → ∞ the minimal λ1 tends to~Nð1Þ= ~Nð2Þ ∼ 1=N, such that F ð2Þ

→ 0.With relatively little effort, one can take the calculation

one step further. Using the first four terms of the Magnusexpansion, the constraint equation (8) can be extended tothird order Hð1Þ

eff þHð3Þeff ¼ H tg with the constraint

Ωtg ¼ ωðA1 þ 2B3 − 4A1A2Þ ð12Þ

defined in terms of Ap ¼P

jfn=ωj2= ~nðpÞ and

Bp ¼P

jfn=ωj4= ~nðpÞ. Equation (10) does not change

to this order since F ð3Þ ¼ 0. The optimal Fourier compo-nents ffng can still be chosen as real and solve the coupledsystem of equations (n ¼ 1;…; N)

fnω

1þ 4A1λ1

~nð2Þ− λ1

1 − 4A2

~nð1Þ

−4λ1f

3n

~nð3Þω3¼

λ2

~nð1Þ: ð13Þ

The full minimization in the third order of expansion,given by the system of Eqs. (13), (9) and (12), can bestraightforwardly solved using the exact second-ordersolution equation (11) as an initial condition for a numeri-cal routine.A neat feature of the optimal pulse is that its total average

intensity IN ¼P

Nn¼1

jfnj2 never exceeds the intensity I1 of

the MC pulse. Since ~nðpÞ−1 decreases with n, the right-hand side of (8) is lower bounded by IN=ω ~Nð1Þ. FromEq. (7) we, however, get I1 ¼ Ωtgω

~Nð1Þ so IN ≤ I1 for thesecond-order optimized pulse. Numerical evidence for thethird-order optimized pulse confirms the relation IN ≤ I1 ina broad range of N.The suppression of fluctuations and deviations from

desired dynamics polychromatic (PC) driving can be see inFig. 1, where populations are depicted over half a periodπ=Ωtg for a pulse with N ¼ 10 frequencies. First of all,panels (a) and (b) of Fig. 1 depict the population of a low-lying state for short and long times, respectively. While theMC evolution with the Rabi frequency, Eq. (7), deviatessignificantly from the target evolution after several periods,the optimal PC dynamics follows the target rather

faithfully. As the amplitude of the MC pulse hasbeen chosen in first order, one might wonder if a betterperformance can be realized with an effective Hamiltonianthat includes higher orders. Such a construction, however,would require a higher driving amplitude and, since theundesired terms in H

ð2Þeff cannot be set to zero, it results in

larger overall deviations with respect to the target dynam-ics. Thus, an improvement of the MC case is not possiblethrough a more accurate treatment. The lower panel (c) ofFig. 1 shows the second main advantage attributed to theshort time scale of one driving period, namely, significantlysmaller fluctuations of the optimal dynamics around thetarget dynamics and, in particular, a considerably lowerpopulation of the intermediate (excited) state.In order to show to what extent the fluctuations can be

suppressed, Fig. 2 depicts the deviations of the actualdynamics around the target unitary as a function of thenumber N of Fourier components. Both the analyticestimate (blue) and the numerically exact solution (red)show that already a moderate number of frequency com-ponents permits us to reduce the fluctuations dramatically.The Fourier components of the optimal pulse for N ¼ 10

are shown in the inset. The component f10 of the highestfrequency Δ ¼ 10ω is close to the MC solution (7); lowerfrequency components are phase shifted by π and theirmagnitude decays rapidly with decreasing frequency.These results show that modulating the driving with onlyfew frequencies suffices to simulate the desired targetunitary with significantly higher precision than in theMC case.The long-time deviations observed in Fig. 1(b) can be

quantitatively measured by the target functional F n, asdefined in Eq. (2), but integrated over the nth drivingperiod. In Fig. 3, these fidelities are shown for the MC pulse

t(a)

(c)

(b)

1 2 3 4 5 6t

0.005

0.010

0.015

0.020

P3

10 20 30 40 50 60

0.2

0.4

0.6

0.8

1.0P2

t

P2

640 650 660 670 680 690

0.2

0.4

0.6

0.8

1.0

FIG. 1. (a) and (b) Transition probability P2ðtÞ ¼jh2jUðtÞj1ij2 for the target dynamics (dotted line), MC dynamics(black line), and third-order optimal dynamics with N ¼ 10

frequency components (gray line). (c) Population of the excitedstate P3 ¼ jh3jUðtÞj1ij2 as a function of time. Plot parameters areΩtg ¼ 0.05ω and n0 ¼ 4.

PRL 113, 010501 (2014) P HY S I CA L R EV I EW LE T T ER Sweek ending4 JULY 2014

010501-3

Page 4: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

and the second- and third-order PC pulses withN ¼ 10 as afunction of n. Let us first compare the MC with the second-order PC pulse. In the first few periods, the optimizedsolution indeed yields a better result. However, the devia-tions with respect to the target dynamics grow faster in thesecond-order optimized case than in the MC case. As aconsequence, in the long-time regime the MC drivingperforms better than the optimized solution calculated insecond order: since the PC pulse contains slower frequen-cies than the MC one, the expansion at second order leadsto a worse approximation of the effective Hamiltonianand the deviations betweenUeff andUtg accumulate in timeand overcome the difference in fluctuations after suffi-ciently long times. Indeed, we observe that the larger

N is, the later the crossover occurs, since the fluctuationsare smaller and deviations from the target Hamiltonianneed more time to accumulate to the value of the MCdynamics. Nevertheless, the optimized PC pulse can alwaysbe systematically improved by pushing the calculations tohigher order in the expansion parameter. As seen inFig. 1 as well as Fig. 3, the third-order optimal pulsesignificantly outperforms the MC dynamics in the entiretime domain.Finally, in order to estimate the robustness of optimal

pulses in realistic experimental setups, we investigate howsmall perturbations to the Fourier components affect theperformance of the optimal pulses. Consider perturbationsof the form

fn → ~fn ¼ fn þ δf; ð14Þ

where δf is a random number uniformly distributed inthe interval [−δ, δ], which accounts for the experimentaluncertainty in the tuned Fourier components. Comparisonbetween the second- and third-order optimal pulsesshows that their largest optimal Fourier componentsdiffer typically by 0.01ω (for Ωtg ¼ 0.05ω), whichdefines a scale δmax for the maximum allowed uncertainty.Perturbations with δ ¼ δmax=4, however, still lead to agood performance; see Fig. 3. Thus, the optimal pulsesappear robust under such perturbations, which indicates agood experimental viability. Moreover, the perturbed,optimized PC pulse provides a significantly lower levelof deviations in comparison with the unperturbed MC pulseeven after several periods 2π=Ωtg.The control of periodically driven systems by means of

pulse shaping presented here opens new perspectives forthe optimal simulation of quantum systems. No increase inintensity as compared to monochromatic driving is requiredand the realization of optimal effective Hamiltonians isrobust under perturbations. Let us point out that thescalings of all parameters are algebraic, so that there isno exponential cost hidden in time, power, or number ofFourier components. The optimal pulses have a rathernarrow spectral range, which eases the identification ofdriving parameters ensuring that no high lying states areexcited. This is of particular importance for large many-body systems, like trapped atomic gases, where uncarefuldriving easily results in uncontrolled heating. Since this canbe avoided with the present approach, it may, for example,be used to enhance or suppress long-range or density-dependent tunneling processes [28] in shaken opticallattices.

Financial support by the European Research Councilwithin the project ODYCQUENT is gratefully acknowl-edged. F. M. acknowledges hospitality by the Centre forQuantum Technologies, a Research Centre of Excellencefunded by the Ministry of Education and the NationalResearch Foundation of Singapore.

0 5 10 15 200.00

0.05

0.10

0.15

Fourier components N

Dev

iati

on

s

2 4 6 8 100.6

0.4

0.2

0.0

0.2

n

fn

FIG. 2 (color online). Magnitude of deviations around the targetunitary dynamics as function of the number N of Fouriercomponents. Blue dots depict the third-order target functionalquantified by Eq. (10) and red triangles depict the numericallyexact target functional, Eq. (2); plot parameters are Ωtg ¼ 0.05ωand n0 ¼ 4. Inset: Fourier components for N ¼ 10.

monochromatic

second order

third order

perturbation third order

0 10 20 30 40 50n

0.05

0.10

0.15

0.20

0.25

0.30n

FIG. 3 (color online). Deviations from the target dynamics F n

in the nth driving period [time interval [ðn − 1ÞT, nT]] for theMC pulse (blue circles), the second-order optimized pulse (brownsquares), and the third-order optimized pulse (red rhombi) withN ¼ 10. The perturbed third-order solution [see Eq. (14)] withfrequency randomness δ ¼ δmax=4, averaged over 100 realiza-tions (green triangles) shows good resilience against experimen-tal uncertainty. Ωtg ¼ 0.05ω and n0 ¼ 4.

PRL 113, 010501 (2014) P HY S I CA L R EV I EW LE T T ER Sweek ending4 JULY 2014

010501-4

Page 5: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

[1] E. Tang, J.-W. Mei, and X.-G. Wen, Phys. Rev. Lett. 106,236802 (2011).

[2] P. Hauke, O. Tieleman, A. Celi, C. Ölschläger, J. Simonet,J. Struck, M. Weinberg, P. Windpassinger, K. Sengstock,M. Lewenstein, and A. Eckardt, Phys. Rev. Lett. 109,145301 (2012).

[3] D. Banerjee, M. Bögli, M. Dalmonte, E. Rico, P. Stebler,U.-J.Wiese, andP. Zoller, Phys.Rev. Lett.110, 125303 (2013).

[4] N. Goldman, A. Kubasiak, A. Bermudez, P. Gaspard, M.Lewenstein, and M. A. Martin-Delgado, Phys. Rev. Lett.103, 035301 (2009).

[5] J. A. Franco-Villafañe, E. Sadurní, S. Barkhofen, U. Kuhl,F. Mortessagne, and T. H. Seligman, Phys. Rev. Lett. 111,170405 (2013).

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

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

[8] H. Miyake, G. A. Siviloglou, C. J. Kennedy, W. C. Burton,and W. Ketterle, Phys. Rev. Lett. 111, 185302 (2013).

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

[10] A. Eckardt, C. Weiss, and M. Holthaus, Phys. Rev. Lett. 95,260404 (2005).

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

[12] J. Struck, C. Ölschläger, M. Weinberg, P. Hauke, J. Simonet,A. Eckardt, M. Lewenstein, K. Sengstock, and P. Wind-passinger, Phys. Rev. Lett. 108, 225304 (2012).

[13] A. Rapp, X. Deng, and L. Santos, Phys. Rev. Lett. 109,203005 (2012).

[14] T. Iadecola, D. Campbell, C. Chamon, C.-Y. Hou, R.Jackiw, S.-Y. Pi, and S. V. Kusminskiy, Phys. Rev. Lett.110, 176603 (2013).

[15] S. Braun, J. P. Ronzheimer, M. Schreiber, S. S. Hodgman,T. Rom, I. Bloch, and U. Schneider, Science 339, 52(2013).

[16] J. Struck, M. Weinberg, C. Olschlager, P. Windpassinger, J.Simonet, K. Sengstock, R. Hoppner, P. Hauke, A. Eckardt,M. Lewenstein, and L. Mathey, Nat. Phys. 9, 738 (2013).

[17] D. D’Alessandro, Introduction to Quantum Control and

Dynamics (CRC, Boca Raton, 2007).[18] V. F. Krotov, Global Methods in Optimal Control Theory

(Marcel Dekker, New York, 1996).[19] N. Khaneja, T. Reiss, C. Kehlet, T. Schulte-Herbrüggen, and

S. J. Glaser, J. Magn. Reson. 172, 296 (2005).[20] B. Rogers, M. Paternostro, G. M. Palma, and G. De Chiara,

Phys. Rev. A 86, 042323 (2012).[21] D. M. Reich, G. Gualdi, and C. P. Koch, Phys. Rev. Lett.

111, 200401 (2013).[22] B. Bartels and F. Mintert, Phys. Rev. A 88, 052315

(2013).[23] G. Floquet, Ann. École Norm. Sup. 12, 47 (1883).[24] See the Supplemental Material at http://link.aps.org/

supplemental/10.1103/PhysRevLett.113.010501 for thetechnical aspects.

[25] W. Magnus, Commun. Pure Appl. Math. 7, 649 (1954).[26] S. Blanes, F. Casas, J. Oteo, and J. Ros, Phys. Rep. 470, 151

(2009).[27] These terms also generate light-shift displacements of the

excited and ground states which do not influence the targetdynamics between the two ground states.

[28] A. Verdeny, A. Mielke, and F. Mintert, Phys. Rev. Lett. 111,175301 (2013).

PRL 113, 010501 (2014) P HY S I CA L R EV I EW LE T T ER Sweek ending4 JULY 2014

010501-5

Page 6: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

Supplemental Material for Optimal Control of Effective Hamiltonians

Albert Verdeny,1, 2 Lukasz Rudnicki,1, 3 Cord A. Muller,4, 5 and Florian Mintert1, 2

1Freiburg Institute for Advanced Studies, Albert-Ludwigs-Universitat, Albertstrasse 19, 79104 Freiburg, Germany2Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom

3Center for Theoretical Physics, Polish Academy of Sciences, Aleja Lotnikow 32/46, 02-668 Warsaw, Poland4Centre for Quantum Technologies, National University of Singapore, Singapore 117543, Singapore

5Department of Physics, University of Konstanz, 78457 Konstanz, Germany(Dated: May 9, 2014)

In this Supplemental Material, several technical as-pects of the analytic minimization are given in detail inorder to ease the reproduction of our results. Accordingto our scheme (i-iii) in the main paper, the target func-tional (TF) Eq. (2) and the constraints Eq. (8), Eq. (9)are the starting point for the minimization . While theconstraints are straightforwardly derived using the Mag-nus expansion, the expressions for the TF require a moredetailed explanation.

This Supplemental Material is divided into three sec-tions:

- In Sec. I, some general aspects regarding the cal-culation of the target functional are explained.

- In Sec. II, the derivation of the TF of the Lambdasystem expanded up to third order

3∑

k=0

F (k) =4

ω2

N∑

n=1

|fn|2

n(2)(s1)

is given in detail.

- In Sec. III, the expression

F (2) = 4λ1Ωtg

ω(s2)

that yields the second order global minimum is de-rived.

I. GENERAL TARGET FUNCTIONAL

Given the decomposition U(t) = U(t)Ueff(t), the TFdefined in Eq. (2) can be written as

F =1

T

∫ T

0

Tr[

2− UUeffU†tg − UtgU

†eff U

†]

dt. (s3)

Under the constraint Htg = Heff , the TF simplifies to

F =1

T

∫ T

0

Tr[

2− U − U†]

dt

=2

T

∫ T

0

(

d−d∑

i

cos (νi(t))

)

dt, (s4)

where d is the dimension of the Hilbert space and νi(t)

are the eigenvalues of G(t) defined via U(t) = e−iG(t).

If the constraint is only satisfied up to a certain order

r, Htg =∑r

k=0 H(k)eff + O(Ωdǫ

r+1), Eq. (s4) is to be

taken up to the same order, F =∑r

k=0 F(k) + O(ǫr+1).

Similarly as with the effective Hamiltonian, an expansionG(t) =

∑rk=1 G

(k)(t) + O(ǫr+1) can be obtained withthe Magnus expansion and the Baker-Cambell-Hausdorffformula. The first two terms read:

G(1)(t) = M1(t) −H(0)eff t, (s5)

G(2)(t) = M2(t) −H(1)eff t +

i

2[G(1)(t), H

(0)eff t]. (s6)

II. TARGET FUNCTIONAL FOR THE

LAMBDA SYSTEM

For the Lambda system described by Eq. (5), the firsttwo leading terms of G(t) are

G(1)(t) = g1(|1〉〈3| + |2〉〈3|) + H.c., (s7)

G(2)(t) = g2 ((|1〉 + |2〉)(〈1| + 〈2|) − 2|3〉〈3|) , (s8)

where

g1 = i∑

n

fnωn

(e−inωt − 1)

+ i∑

n

fnω(n + Nn0)

(e−i(n+Nn0)ωt − 1). (s9)

The second order term g2 will not contribute to the finalresult. The eigenvalues of the sum G(1)(t) + G(2)(t) are

ν1, ν2, ν3 = 0,±√

2(|g1|2 + 2g22). Consistently withour expansion in ǫ, the cosine in Eq. (s4) can be expandedin a Taylor series up to second order in its argument,namely (d = 3)

3∑

k=0

F (k) =2

T

∫ T

0

(

3 −

3∑

i=1

(

1 −ν2i (t)

2

)

)

dt

=2

T

∫ T

0

3∑

i=1

ν2i (t)

2dt. (s10)

As the eigenvalues νi(t) enter quadratically in (s10), theonly contribution to the TF is given by G(1)(t); G(2)(t)and higher orders of the cosine yield contributions offourth or higher order in ǫ. Thus

3∑

k=0

F (k) = F (2) =4

T

∫ T

0

|g1|2dt. (s11)

Page 7: Optimal Control of Effective Hamiltonianstheorie.physik.uni-konstanz.de/burkard/sites/default/files/images/... · Optimal Control of Effective Hamiltonians Albert Verdeny,1,2 Łukasz

2

Using the expression Eq. (s9) and the constraint (9)results in Eq. (s1).

III. MINIMIZATION IN SECOND ORDER

A crucial point for the analytic results obtained in sec-ond order is the possibility to write the target functionalin terms of the Lagrange multiplier λ1 as in Eq. (s2),which permits us to obtain the global minimum of F (2).

The TF to be minimized and the two constraints

C1 =∑

n

|fn|2

ω2n(1)−

Ωtg

ωand C2 =

n

f∗n

ωn(1)(s12)

define the Lagrange functional

L[fn, f∗n, λ1, λ2] = F (2)/4 − λ1C1 − λ2C2. (s13)

The extrema of L are given by the system of Eqs. (8), (9)and (11), which are found after differentiating (s13) withrespect to λ1, λ2 and f∗

n respectively. After insertingEq. (11) into Eqs. (8) and (9), two equations for theLagrange multipliers are obtained

0 =

N∑

n=1

(

n(1)2

n(2)− λ1n(1)

)−1

, (s14)

|λ2| =

Ωtg

ω

(

N∑

n=1

n(1)

(

n(1)2

n(2)− λ1n(1)

)−2)−1/2

.(s15)

Eq. (s15) can be rewritten using (s14) as

|λ2| =

Ωtg

ω

(

N∑

n=1

(

n(1) −n(1)2

λ1n(2)+

n(1)2

λ1n(2)

)(

n(1)2

n(2)− λ1n(1)

)−2)−1/2

=

λ1Ωtg

ω

(

N∑

n=1

n(1)2

n(2)

(

n(1)2

n(2)− λ1n(1)

)−2)−1/2

. (s16)

Finally, inserting Eq. (11) into Eq. (10) and using Eq. (s16) above results in Eq. (s2).