shortcuts to dynamic polarizationshortcuts to dynamic polarization tamiro villazon, 1pieter w....

22
Shortcuts to Dynamic Polarization Tamiro Villazon, 1 Pieter W. Claeys, 2, * Anatoli Polkovnikov, 1 and Anushya Chandran 1 1 Department of Physics, Boston University, 590 Commonwealth Ave., Boston, MA 02215, USA 2 TCM Group, Cavendish Laboratory, University of Cambridge, Cambridge CB3 0HE, UK Dynamic polarization protocols aim to hyperpolarize a spin bath by transferring spin polarization from a well- controlled qubit such as a quantum dot or a color defect. Building on techniques from shortcuts to adiabaticity, we design fast and ef๏ฌcient dynamic polarization protocols in central spin models that apply to dipolarly interacting systems. The protocols maximize the transfer of polarization via bright states at a nearby integrable point, exploit the integrability-breaking terms to reduce the statistical weight on dark states that do not transfer polarization, and realize experimentally accessible local counterdiabatic driving through Floquet-engineering. A master equation treatment suggests that the protocol duration scales linearly with the number of bath spins with a pre-factor that can be orders of magnitude smaller than that of unassisted protocols. This work opens new pathways to cool spin baths and extend qubit coherence times for applications in quantum information processing and metrology. I. INTRODUCTION A prevalent goal in several ๏ฌelds of physics and chemistry is to ef๏ฌciently polarize an ensemble of spin particles. In nuclear magnetic resonance spectroscopy (NMR) and mag- netic resonance imaging (MRI), polarizing nuclear spins en- hances sensitivity and resolution [1โ€“4]. In applications to quantum information processing, hyperpolarization schemes can be used to initialize large-scale quantum simulators [5] or to extend qubit coherence times by cooling the surrounding spin bath [6, 7]. Where costly or dif๏ฌcult to polarize the spin ensemble directly, dynamic polarization protocols have been developed to repeatedly transfer polarization from readily po- larized control spins [1, 8โ€“14]. In simple experimental setups, a spin bath is polarized by controlling a single qubit, such as a nitrogen vacancy (NV) center in diamond [15โ€“17] or a quantum dot [18โ€“20], whose polarization can be repeatedly reset, effectively generating a zero temperature reservoir for the bath [21]. A key goal of this article is to introduce a fast and ef๏ฌcient scheme for dynamic polarization in central spin models. Polarization transfer relies on the spin-๏ฌ‚ip interactions be- tween a control spin and the spin ensemble to be polarized. The Hamiltonian can be schematically represented as =ฮฉ( ) + spin-๏ฌ‚ip , (1) consisting of an electromagnetic ๏ฌeld ฮฉ( ) acting on the con- trol spin along the z-direction and spin-๏ฌ‚ip interactions be- tween control spin and spin bath. Given an initially polarized control spin, ฮฉ( ) can be tuned to transfer polarization [22โ€“24]. Speci๏ฌcally, dynamic polarization protocols can be separated into two classes: (i) sudden protocols in which the control ๏ฌeld ฮฉ( ) is quenched to resonance with the spin-๏ฌ‚ip interactions to induce polarization transfer and (ii) adiabatic protocols in which polarization transfer is induced by slowly driving ฮฉ( ) across resonances [25]. Adiabatic protocols offer an advantage over sudden protocols as they do not require precise resonance tuning and pulse timing. They also can cover a broader range of * [email protected] bath spin resonances, enabling robust transfer in the presence of ๏ฌeld and interaction inhomogeneities [26โ€“28]. Their main disadvantage is the requirement of slow speeds, which can be inef๏ฌcient or unfeasible in experiments limited by spin diffu- sion in the bath and decoherence of the control spins [29, 30]. Apart from the limitations on control speeds, the achievable polarization is also limited by the presence of dark states, mak- ing it seldom possible to completely polarize the bath even at slow speeds [21, 31โ€“34]. Dark states are many-body qubit- bath eigenstates in which the qubit is effectively decoupled from the bath. Since such states have a ๏ฌxed control spin po- larization and cannot be depopulated through changes in the qubit control ๏ฌeld, any initial nonzero population of dark states will limit hyperpolarization. Experiments in different material systems have found maximum saturation at about 60% full polarization [35โ€“37]. Several schemes have been proposed to enhance hyperpo- larization by depopulating dark states effectively [20], for ex- ample by modulating the electron wavefunction of the qubit in quantum dots [21, 31] or by alternating resonant drives which reduce quantum correlations in the bath [38]. While studies so far mainly focused spin systems where the central spin interacts with its environment through fully isotropic (XXX) interac- tions, arising in e.g. quantum dots in semiconductors, we consider a model where the interactions are anisotropic (XX), as in resonant dipolar spin systems [17, 19, 23, 24, 38โ€“40]. Overcoming the requirement of slow speeds in adiabatic protocols is the aim of the ๏ฌeld of shortcuts to adiabaticity [41, 42]. Shortcut methods such as counterdiabatic driving (CD) suppress diabatic transitions between the eigenstates of a driven Hamiltonian ( ) by evolving the system with a Hamil- tonian ( ) containing additional counter terms [43โ€“49]. CD preserves the systemโ€™s adiabatic path through state space even during ultra-fast protocols. CD protocols typically require engineering operators which are highly complex and many- body, making them dif๏ฌcult to implement in practice [49]. Recent progress has focused on reducing the complexity of CD Hamiltonians, for example by mapping them to simpler unitary equivalents [50โ€“53], or by approximating them with lo- cal (few-body) operators [54โ€“56]. The required local operators can be realized through e.g. Floquet-engineering techniques, using high-frequency oscillations to realize the CD Hamilto- arXiv:2011.05349v2 [quant-ph] 16 Feb 2021

Upload: others

Post on 04-Feb-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

  • Shortcuts to Dynamic Polarization

    Tamiro Villazon,1 Pieter W. Claeys,2, โˆ— Anatoli Polkovnikov,1 and Anushya Chandran11Department of Physics, Boston University, 590 Commonwealth Ave., Boston, MA 02215, USA

    2TCM Group, Cavendish Laboratory, University of Cambridge, Cambridge CB3 0HE, UK

    Dynamic polarization protocols aim to hyperpolarize a spin bath by transferring spin polarization from a well-controlled qubit such as a quantum dot or a color defect. Building on techniques from shortcuts to adiabaticity, wedesign fast and efficient dynamic polarization protocols in central spin models that apply to dipolarly interactingsystems. The protocols maximize the transfer of polarization via bright states at a nearby integrable point, exploitthe integrability-breaking terms to reduce the statistical weight on dark states that do not transfer polarization, andrealize experimentally accessible local counterdiabatic driving through Floquet-engineering. A master equationtreatment suggests that the protocol duration scales linearly with the number of bath spins with a pre-factor thatcan be orders of magnitude smaller than that of unassisted protocols. This work opens new pathways to cool spinbaths and extend qubit coherence times for applications in quantum information processing and metrology.

    I. INTRODUCTION

    A prevalent goal in several fields of physics and chemistryis to efficiently polarize an ensemble of spin particles. Innuclear magnetic resonance spectroscopy (NMR) and mag-netic resonance imaging (MRI), polarizing nuclear spins en-hances sensitivity and resolution [1โ€“4]. In applications toquantum information processing, hyperpolarization schemescan be used to initialize large-scale quantum simulators [5] orto extend qubit coherence times by cooling the surroundingspin bath [6, 7]. Where costly or difficult to polarize the spinensemble directly, dynamic polarization protocols have beendeveloped to repeatedly transfer polarization from readily po-larized control spins [1, 8โ€“14]. In simple experimental setups,a spin bath is polarized by controlling a single qubit, suchas a nitrogen vacancy (NV) center in diamond [15โ€“17] or aquantum dot [18โ€“20], whose polarization can be repeatedlyreset, effectively generating a zero temperature reservoir forthe bath [21]. A key goal of this article is to introduce a fastand efficient scheme for dynamic polarization in central spinmodels.

    Polarization transfer relies on the spin-flip interactions be-tween a control spin and the spin ensemble to be polarized.The Hamiltonian can be schematically represented as

    ๐ป = ฮฉ(๐‘ก) ๐‘†๐‘ง + ๐ปspin-flip , (1)

    consisting of an electromagnetic field ฮฉ(๐‘ก) acting on the con-trol spin along the z-direction and spin-flip interactions be-tween control spin and spin bath. Given an initially polarizedcontrol spin, ฮฉ(๐‘ก) can be tuned to transfer polarization [22โ€“24].Specifically, dynamic polarization protocols can be separatedinto two classes: (i) sudden protocols in which the control fieldฮฉ(๐‘ก) is quenched to resonance with the spin-flip interactionsto induce polarization transfer and (ii) adiabatic protocols inwhich polarization transfer is induced by slowly driving ฮฉ(๐‘ก)across resonances [25]. Adiabatic protocols offer an advantageover sudden protocols as they do not require precise resonancetuning and pulse timing. They also can cover a broader range of

    โˆ— [email protected]

    bath spin resonances, enabling robust transfer in the presenceof field and interaction inhomogeneities [26โ€“28]. Their maindisadvantage is the requirement of slow speeds, which can beinefficient or unfeasible in experiments limited by spin diffu-sion in the bath and decoherence of the control spins [29, 30].

    Apart from the limitations on control speeds, the achievablepolarization is also limited by the presence of dark states, mak-ing it seldom possible to completely polarize the bath evenat slow speeds [21, 31โ€“34]. Dark states are many-body qubit-bath eigenstates in which the qubit is effectively decoupledfrom the bath. Since such states have a fixed control spin po-larization and cannot be depopulated through changes in thequbit control field, any initial nonzero population of dark stateswill limit hyperpolarization. Experiments in different materialsystems have found maximum saturation at about 60% fullpolarization [35โ€“37].

    Several schemes have been proposed to enhance hyperpo-larization by depopulating dark states effectively [20], for ex-ample by modulating the electron wavefunction of the qubit inquantum dots [21, 31] or by alternating resonant drives whichreduce quantum correlations in the bath [38]. While studies sofar mainly focused spin systems where the central spin interactswith its environment through fully isotropic (XXX) interac-tions, arising in e.g. quantum dots in semiconductors, weconsider a model where the interactions are anisotropic (XX),as in resonant dipolar spin systems [17, 19, 23, 24, 38โ€“40].

    Overcoming the requirement of slow speeds in adiabaticprotocols is the aim of the field of shortcuts to adiabaticity[41, 42]. Shortcut methods such as counterdiabatic driving(CD) suppress diabatic transitions between the eigenstates of adriven Hamiltonian ๐ป (๐‘ก) by evolving the system with a Hamil-tonian ๐ป๐ถ๐ท (๐‘ก) containing additional counter terms [43โ€“49].CD preserves the systemโ€™s adiabatic path through state spaceeven during ultra-fast protocols. CD protocols typically requireengineering operators which are highly complex and many-body, making them difficult to implement in practice [49].Recent progress has focused on reducing the complexity ofCD Hamiltonians, for example by mapping them to simplerunitary equivalents [50โ€“53], or by approximating them with lo-cal (few-body) operators [54โ€“56]. The required local operatorscan be realized through e.g. Floquet-engineering techniques,using high-frequency oscillations to realize the CD Hamilto-

    arX

    iv:2

    011.

    0534

    9v2

    [qu

    ant-

    ph]

    16

    Feb

    2021

    mailto:[email protected]

  • 2

    nian as an effective high-frequency Hamiltonian using onlycontrols present in the original adiabatic protocol [55, 57โ€“60].Local counterdiabatic driving has recently been realized ex-perimentally in synthetic tight-binding lattices [61], in IBMโ€™ssuperconducting quantum computer [62], and in a liquid-stateNMR system for a nonintegrable spin chain [63]. Such meth-ods have also gained attention in the context of quantum ther-modynamics, where (approximate) CD can be used to speed upunderlying adiabatic processes and increase the performanceof quantum engines [58, 64โ€“71].

    We develop a dynamic polarization scheme which imple-ments approximate counterdiabatic driving (CD) to quicklyand efficiently polarize a spin bath using a tunable qubit whilesimultaneously depopulating dark states. In the absence ofinhomogeneous bath fields, the model Hamiltoniani is inte-grable [34]. We first exploit the integrability of this modelto design a CD protocol explicitly targeting all polarization-transferring bright (i.e., not dark) states. Within all protocolsthe bright states arise in pairs acting as independent two-levelLandau-Zener systems, for which CD protocols can be straight-forwardly designed. While the exact CD protocol targets allbright states and gives rise to a highly involved control Hamilto-nian, we show how the CD protocol can be well approximatedusing local (few-body) operators and experimentally imple-mented using Floquet engineering (FE). In the presence ofinhomogeneous bath fields the system is no longer integrable.However, the proposed protocols still lead to a remarkableincrease in transfer efficiency. Furthermore, the local counter-diabatic (LCD) protocol dynamically couples dark states tobright states, such that dark states can be depopulated. Notonly are such LCD protocols much easier to implement thanthe exact CD ones, we find that they outperform CD protocolsand lead to a complete hyperpolarization of the spin bath.

    The FE protocols also lead to natural quantum speed limits:there exists a lower bound for the protocol durations belowwhich the FE protocol can no longer accurately mimic theLCD protocol. The emergence of speed limits is ubiquitousin shortcut protocols and control theory [42, 49, 52, 71โ€“78].Interestingly, our work now suggests that speed limits are alsointrinsic in approximate local counterdiabatic protocols.

    This paper is organized as follows. In Section II, we presentthe qubit-bath model system and the hyperpolarization schemeused throughout this work. In Section III, we construct anddetail the CD and LCD protocols and compare their efficiencywith unassisted (UA) protocols which do not use shortcut meth-ods. In Section IV, we show how our shortcut protocols canbe applied to fully polarize a spin bath. A master equationfor the hyperpolarization is introduced in Section V, which isused to analyze the protocols at large system sizes and showthat all protocol durations scale linearly with the number ofbath spins. In Section VI, we show how to realize LCD withFE and discuss the emergence of a quantum speed limit. Weconclude in Section VII with a discussion of our results in abroader context.

    II. MODEL AND HYPERPOLARIZATION SCHEME

    A. Hamiltonian

    We focus on a concrete central spin model describing a qubitinteracting with ๐ฟ โˆ’ 1 spin-1/2 bath spins. The Hamiltonian isgiven by

    ๐ป (๐‘ก) = ฮฉ๐‘„ (๐‘ก) ๐‘†๐‘ง0+๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    ฮฉ๐ต, ๐‘— ๐‘†๐‘ง๐‘—+ 1

    2

    ๐ฟโˆ’1โˆ‘๏ธ๐‘—=1๐‘” ๐‘—

    (๐‘†+0๐‘†

    โˆ’๐‘— +๐‘†โˆ’0 ๐‘†

    +๐‘—

    ), (2)

    where ฮฉ๐‘„ (๐‘ก) is the magnetic field strength on the qubit, ฮฉ๐ต, ๐‘—is the magnetic field strength on the ๐‘— th bath spin, and ๐‘” ๐‘— is thecoupling strength between the qubit and the ๐‘— th bath spin, with๐‘— = 1, 2, . . . , ๐ฟ โˆ’1. Eq. (2) describes several physical setups inrotating frames, such as color defects or quantum dots coupledto ensembles of nuclear spins via dipolar interactions [17, 19,23, 24, 38]. Spin conserving (โ€˜flip-flopโ€™) transitions dominatethe dipolar interaction provided ๐‘” ๐‘— ๏ฟฝ ฮฉ๐‘„ + ฮฉ๐ต, with thelatter set by the amplitudes of the continuous driving fields; astandard derivation is given in Appendix A. The top panel ofFig. 1 shows a schematic of the model.

    Experimentally, the bath field and qubit-bath couplings arespatially inhomogeneous. For simplicity, we model these in-homogeneities as uncorrelated disorder: we draw each ฮฉ๐ต, ๐‘—independently from a uniform distribution

    ฮฉ๐ต, ๐‘— โˆˆ [ฮฉ๐ต โˆ’ ๐›พ๐‘ง ,ฮฉ๐ต + ๐›พ๐‘ง], (3)

    where ฮฉ๐ต sets the mean value and ๐›พ๐‘ง sets the z-disorderstrength. We also draw each ๐‘” ๐‘— independently from a uniformdistribution

    ๐‘” ๐‘— โˆˆ [๐‘” โˆ’ ๐›พ๐‘ฅ๐‘ฅ , ๐‘” + ๐›พ๐‘ฅ๐‘ฅ], (4)

    where ๐‘” sets the mean value and ๐›พ๐‘ฅ๐‘ฅ sets the xx-disorderstrength. In this work, we probe the weak coupling and disor-der regime given by ๐›พ๐‘ฅ๐‘ฅ , ๐›พ๐‘ง < ๐‘” ๏ฟฝ ฮฉ๐ต.

    Since ๐ป conserves total magnetization [๐ป, ๐‘€] = 0, where

    ๐‘€ โ‰ก๐ฟโˆ’1โˆ‘๏ธ๐‘—=0

    ๐‘†๐‘ง๐‘—, (5)

    its eigenspectrum splits into ๐ฟ + 1 polarization sectors (seeleft of lower panel in Fig 1). Each sector can alternativelybe specified by the number ๐‘ = ๐‘€ + ๐ฟ/2 of spin flips abovethe fully-polarized state |โ†“ใ€‰ โŠ— |โ†“โ†“ . . . โ†“ใ€‰. The aim of hyperpo-larization is then to find protocols that systematically reduce๐‘€, where a fully polarized state corresponds to ๐‘ = 0 or๐‘€ = โˆ’๐ฟ/2.

    B. Spectrum

    The eigenstates of ๐ป capture essential features common inapplications of dynamic polarization: bright states which allowresonant polarization transfer when ฮฉ๐‘„ (๐‘ก) is varied, and darkstates which limit transfer. In the ๐›พ๐‘ง = 0 limit, the model is

  • 3

    FIG. 1. Model schematic, spectrum, and hyperpolarizationscheme. (Top panel) Schematic of the central spin model in Eq. (2).(Bottom panel) On the left, we illustrate the polarization sectors andspectrum for a system with an even number ๐ฟ of spins and smalldisorder strengths. On the right, we illustrate the spectrum in twopolarization sectors with ๐‘€ < 0, ฮ”max sets the maximal width of theresonance region, where together with an outline of the transfer-resethyperpolarization scheme.

    integrable and the structure of its eigenstates is known [34].While our proposed protocols are not restricted to integrablemodels, the known eigenstate structure at the integrable pointallows for a quantitative understanding of the general coolingprotocols. We briefly review these eigenstates and their basicproperties in the integrable limit, and subsequently extend thediscussion to ๐›พ๐‘ง > 0.

    1. Bright States (๐›พ๐‘ง = 0)

    Bright eigenstates can be written as

    |B๐›ผ (๐œ†)ใ€‰ = ๐‘๐›ผโ†‘ (๐œ†) |โ†‘ใ€‰ โŠ— |B๐›ผโ†‘ ใ€‰ + ๐‘

    ๐›ผโ†“ (๐œ†) |โ†“ใ€‰ โŠ— |B

    ๐›ผโ†“ ใ€‰ , (6)

    where

    ๐œ†(๐‘ก) โ‰ก ฮฉ๐‘„ (๐‘ก) โˆ’ฮฉ๐ต (7)

    measures the detuning between the qubit and bath fields. Onresonance, ฮฉ๐‘„ = ฮฉ๐ต and ๐œ† = 0. The index ๐›ผ distinguishesbetween the different bright states. Crucially, the bath states|B๐›ผโ†‘,โ†“ใ€‰ do not depend on ๐œ†.

    As such, when varying ๐œ† the bright states only couple inpairs (๐›ผ = ยฑ๐‘˜), behaving as independent two-level Landau-Zener systems. The Hamiltonian in each such two-dimensionalsubspace can be written as

    ๐ป๐›ผ (๐œ†) = ๐œ†๐‘†๐‘ง๐›ผ + ฮ”๐›ผ๐‘†๐‘ฅ๐›ผ +ฮฉ๐ต ๐‘€, (8)

    where ฮ”๐›ผ sets the energy splitting (gap) of the pair at resonance[34] and we have introduced generalized spin operators ๐‘†๐‘ฅ,๐‘ฆ,๐‘ง๐›ผacting on the two-dimensional space spanned by |B๐›ผ+ ใ€‰ = |โ†‘ใ€‰ โŠ—|B๐›ผโ†‘ ใ€‰ and |B

    ๐›ผโˆ’ ใ€‰ = |โ†“ใ€‰ โŠ— |B๐›ผโ†“ ใ€‰.

    ๐‘†๐‘ฅ๐›ผ =12

    (|B๐›ผโˆ’ ใ€‰ ใ€ˆB๐›ผ+ | + |B๐›ผ+ ใ€‰ ใ€ˆB๐›ผโˆ’ |

    ),

    ๐‘†๐‘ฆ๐›ผ =

    ๐‘–

    2(|B๐›ผโˆ’ ใ€‰ ใ€ˆB๐›ผ+ | โˆ’ |B๐›ผ+ ใ€‰ ใ€ˆB๐›ผโˆ’ |

    ),

    ๐‘†๐‘ง๐›ผ =12

    (|B๐›ผ+ ใ€‰ ใ€ˆB๐›ผ+ | โˆ’ |B๐›ผโˆ’ ใ€‰ ใ€ˆB๐›ผโˆ’ |

    ).

    ๐‘†๐‘ง๐›ผ corresponds to ๐‘†๐‘ง0 projected on a bright pair subspace. The

    apparent simplicity of the problem in this subspace hides thecomplexity of the qubit-bath interactions present in the orig-inal spin basis, where the bath states |B๐›ผโ†‘,โ†“ใ€‰ and the gap ฮ”๐›ผare obtained by solving a set of nonlinear Bethe equations[34]. Within each magnetization sector ๐‘€ = ๐‘ โˆ’ ๐ฟ/2, we la-bel bright state pairs by ๐›ผ = |๐‘˜ |, where ๐‘˜ โˆˆ {1, 2, . . . , ๐‘›๐ต}and

    ๐‘›๐ต =

    (๐ฟ โˆ’ 1๐‘ โˆ’ 1

    ), (9)

    is the number of pairs in the sector for ๐‘€ < 0 1.The Hamiltonian (8) returns the bright state energies

    ๐ธ ๐›ผB (๐œ†) = ฮฉ๐ต ๐‘€ ยฑ12

    โˆš๏ธƒ๐œ†2 + ฮ”2๐›ผ . (10)

    We refer to the set of bright states with positive๐ธ ๐›ผB โˆ’ฮฉ๐ต๐‘€ > 0 as the top bright band, and those with neg-ative ๐ธ ๐›ผB โˆ’ฮฉ๐ต๐‘€ < 0 as the bottom bright band (see red bandsin bottom panel of Fig. 1).

    In bright states, polarization can be transferred between thequbit and the bath. At resonance (๐œ† = 0) the bright state pairsare fully hybridized with ๐‘ยฑ๐‘˜โ†‘ = ยฑ๐‘

    ยฑ๐‘˜โ†“ . Initializing the system

    in a fully polarized central spin state and then quenching toresonance, as is done in sudden protocols, transfers polarizationon the timescale ฮ”โˆ’1๐›ผ . Adiabatic protocols induce a qubit-bathpolarization transfer in bright states by slowly varying ๐œ†(๐‘ก)resonance. As ๐œ† โ†’ ยฑโˆž, bright states approach a product formand the initial eigenstate |โ†‘ใ€‰ โŠ— |B๐›ผโ†‘ ใ€‰ is adiabatically connectedto |โ†“ใ€‰ โŠ— |B๐›ผโ†“ ใ€‰ and vice versa.

    Each Landau-Zener problem is fully characterized by itsgap. The distribution of bright pair gaps at resonance ฮ”๐›ผ isshown in Fig. 2 for various polarization sectors. As shown in

    1 Since the aim of the proposed protocols is to reduce magnetization, wefocus on ๐‘€ < 0.

  • 4

    Appendix B, the gap distribution can be obtained analyticallyat zero disorder (๐›พ๐‘ฅ๐‘ฅ = 0) in the thermodynamic limit wherewe take ๐ฟ โ†’ โˆž, holding ๐‘š โ‰ก |๐‘€ |/๐ฟ fixed. Since these gapsset the necessary time scales for adiabatic protocols, we brieflydetail some relevant gap scales. The typical gap scale is givenby

    ฮ”typ โ‰กโˆš๏ธƒโˆ‘๏ธ

    ๐‘”2๐‘—โˆผโˆš๐ฟ โˆ’ 1 ๐‘”, (11)

    shown as a gold vertical dashed line in Fig. 2, whereas themaximal gap, also setting the maximal width of the resonanceregion, is given by

    ฮ”max โˆผ ๐ฟ ๐‘”, (12)

    scaling extensively in ๐ฟ. The smallest bright gap in the homo-geneous model can be found as

    ฮ”min = ๐‘”โˆš๏ธ

    2(๐‘€ + 1) โˆผ ๐‘”โˆš

    2๐‘š ๐ฟ. (13)

    At sufficiently small gaps ฮ” & ฮ”min, the distribution of brightpair gaps is given by

    ๐‘›(ฮ”) โˆ ฮ”(1 + 2๐‘š1 โˆ’ 2๐‘š

    )โˆ’(ฮ”/ฮ”min)2; ฮ” โ‰ฅ ฮ”min. (14)

    This distribution is shown in Fig. 2 as a dashed black curve.We find good qualitative agreement between the analyticalcurve at ๐›พ๐‘ฅ๐‘ฅ = 0 and our numerical results for small but finitedisorder ๐›พ๐‘ฅ๐‘ฅ = 0.05 in the magnetization sector ๐‘€ = โˆ’1 withthe largest Hilbert space dimension.

    Fig. 2 also shows the numerically obtained distribution ofbright pair gaps in the ๐‘€ = โˆ’4 and ๐‘€ = โˆ’7 sectors. Thedistribution in the ๐‘€ = โˆ’4 sector exhibits three broad peakswhich are centered around the three bright pair gap energiesin the ๐›พ๐‘ฅ๐‘ฅ = 0 limit (Appendix B). As the width of each peakis proportional to ๐›พ๐‘ฅ๐‘ฅ๐ฟ while the bright pair gaps at ๐›พ๐‘ฅ๐‘ฅ = 0are order one, we expect the three-peak structure to be washedout at larger ๐ฟ, and the distribution to be captured by Eq. (14)instead. In the ๐‘€ = โˆ’7 sector, we expect a single pair ofbright states with pair gap โ‰ˆ ฮ”typ at small ๐›พ๐‘ฅ๐‘ฅ , as confirmedby Fig. 2. We note that Eq. (14) only applies to sectors withfinite magnetization density at large ๐ฟ.

    The main difference comes from the non-zero density ofgaps for ฮ” < ฮ”min. However, as will be shown in followingsections, this non-zero density does not qualitatively influenceour protocols.

    In sum, the bright bands consist of an ensemble of indepen-dent Landau-Zener systems with a non-trivial distribution ofgaps. For each bright pair, an adiabatic passage of ๐œ† across res-onance prevents excitations across its gap and flips polarizationof the qubit, transferring polarization to the bath.

    2. Dark States (๐›พ๐‘ง = 0)

    Dark states have the following product form with the centralqubit fully polarized:

    |D๐›ผใ€‰ = |โ†‘ใ€‰ โŠ— |D๐›ผโ†‘ ใ€‰ or |D๐›ผใ€‰ = |โ†“ใ€‰ โŠ— |D๐›ผโ†“ ใ€‰ , (15)

    FIG. 2. Distribution of bright pair gaps at resonance. Histogramof number n(ฮ”๐›ผ) of bright state pairs with gap ฮ”๐›ผ. Data is shown fora typical disorder realization in multiple polarization sectors. The goldvertical dashed line denotes the typical gap ฮ”typ. The black dashedcurve denotes the distribution of gaps from Eq. (14). Parameters:๐ฟ = 16, ๐œ† = 0, ๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0, and 60 bins.

    where the index ๐›ผ distinguishes between the different darkstates. The bath states |D๐›ผโ†‘,โ†“ใ€‰ depend implicitly on {๐‘” ๐‘— }, butcrucially not on ๐œ†, and are obtained by solving a set of โ€˜darkโ€™Bethe equations [34].

    In a given polarization sector ๐‘€ = ๐‘ โˆ’ ๐ฟ/2, there are

    ๐‘›๐ท =

    ๏ฟฝ๏ฟฝ๏ฟฝ๏ฟฝ(๐ฟ โˆ’ 1๐‘ ) โˆ’ (๐ฟ โˆ’ 1๐‘ โˆ’ 1)๏ฟฝ๏ฟฝ๏ฟฝ๏ฟฝ (16)dark states. Dark states with central qubit polarized along +๐‘งonly exist in sectors ๐‘€ > 0, while dark states with central spinpolarization along โˆ’๐‘ง only exist in sectors ๐‘€ < 0, and no darkstates exist in the sector ๐‘€ = 0. Dark state are eigenstates of ๐‘†๐‘ง0and are annihilated by the interaction part of the Hamiltonian[34], such that the energies given by

    ๐ธ ๐›ผD (๐œ†) = ฮฉ๐ต ๐‘€ + sgn[๐‘€]๐œ†

    2, (17)

    change linearly with the qubit field detuning ๐œ†. Their wavefunctions however do not change with ๐œ†, preventing polariza-tion transfer to the bath.

    3. Bright & Dark States (๐›พ๐‘ง > 0)

    In the presence of z-disorder (๐›พ๐‘ง > 0), the system is notintegrable. However, the same qualitative picture for the eigen-states holds: on adiabatically changing the detuning ๐œ† andcomparing the polarization of the central spin far away fromresonance (๐œ† = ยฑโˆž), there exists a subset of โ€˜bright statesโ€™ inwhich the polarization is changed and a subset of โ€˜dark statesโ€™for which the polarization is unchanged.

    Since the central spin is polarized far away from resonance,a counting argument can be used to determine the number ofbright and dark states. Consider a sector with magnetization

  • 5

    ๐‘€ < 0 and dimension ๐‘›๐‘€ =( ๐ฟ๐‘

    ): there are ๐‘›โ†“ =

    (๐ฟโˆ’1๐‘

    )states

    in which the qubit is fully polarized along the โˆ’๐‘ง-direction and๐‘›โ†‘ =

    ( ๐ฟโˆ’1๐‘โˆ’1

    )states in which the qubit is fully polarized along

    the +๐‘ง-direction. Far from resonance, the energies are given byฮฉ๐ต๐‘€ โˆ’ ๐œ†/2 and ฮฉ๐ต + ๐œ†/2 respectively. Comparing the totalnumber of states in the top and bottom band far away fromresonance, there must be ๐‘›๐ท = ๐‘›โ†“ โˆ’ ๐‘›โ†‘ dark states in whichthe polarization of the qubit does not flip for an adiabaticpassage across resonance (see also Fig. 1). The remaining๐‘›๐‘€ โˆ’ ๐‘›๐ท = 2 ๐‘›โ†‘ states are bright states in which the spin of thequbit flips during such an adiabatic process, consistent withEqs. (9) and (16). This simple counting argument only usesconservation of total z-magnetization and produces the samequalitative eigenstate band structure as one sthe Bethe ansatzin the integrable limit (๐›พ๐‘ง = 0) [34].

    There are, however, important differences between the in-tegrable (๐›พ๐‘ง = 0) and non-integrable (๐›พ๐‘ง > 0) models in theresonance regime. When ๐›พ๐‘ง > 0, the simple product state struc-ture of dark states and the Landau-Zener structure of brightstates is no longer exact: the non-integrable eigenstates aremixtures of the unperturbed states and exhibit ergodic behavior(Appendix E).

    While adiabatic protocols transferring polarization in theinhomogeneous model are qualitatively similar to those inthe homogeneous model, and bright and dark states can gen-erally be defined by their central spin polarization far awayfrom resonance, non-adiabatic effects can enhance polariza-tion transfer in the inhomogeneous model. Finite z-disorderis useful for the purposes of dynamic polarization: in a non-adiabatic protocol dark states can be excited to bright statessince |ใ€ˆD๐›ผ |๐‘†๐‘ง0 |B

    ๐›ผโ€ฒใ€‰| > 0. Dark states in the inhomogeneousmodel can be depopulated during a non-adiabatic passageacross resonance, such that the limit on hyperpolarization canbe overcome by preferentially inducing transitions from darkstates to bright states.

    C. Hyperpolarization Scheme

    We now discuss the basic hyperpolarization scheme as illus-trated in Fig. 1.

    To polarize the spin bath, we apply a cyclical scheme. Ineach cycle, we (i) reset the polarization of the qubit to |โ†“ใ€‰ atlarge detuning, and (ii) we transfer polarization from the qubitto the bath by sweeping the central field detuning ๐œ†(๐‘ก) acrossresonance over a timescale ๐œ๐‘Ÿ . The reset step is a routine ex-perimental step in quantum computing platforms; for example,in a NV set-up, the qubit can be reset using a rapid opticalpulse [59, 79]. The ramp varies ๐œ†(๐‘ก) from an initial value ๐œ†๐‘–to a final value ๐œ† ๐‘“ = โˆ’๐œ†๐‘– , such that the cycle starts and endsfar from resonance ๐œ†0 โ‰ก |๐œ†๐‘– | = |๐œ† ๐‘“ | ๏ฟฝ ฮ”max, where the qubitis completely polarized in every eigenstate. From one cycle tothe next, the direction of the ramp is reversed (after each reset)in a forward-backward fashion.

    During a single reset and sweep cycle probability is trans-

    ferred in every magnetization sector2 (๐‘€) from states with upqubit polarization |โ†‘ใ€‰ in sector ๐‘€ to states with down qubitpolarization |โ†“ใ€‰ in the magnetization sector (๐‘€ โˆ’ 1) (as de-picted in Fig. 1). The effects on a single bright state can bereadily understood: suppose the system is initially in a brighteigenstate |โ†“ใ€‰ โŠ— |B๐›ผโ†“ ใ€‰, factorizable far away from resonanceand with fixed magnetization ๐‘€ . Then the total magnetizationof the bath state is necessarily ๐‘€ + 1/2. After an ideal adi-abatic transfer across resonance, this bright state is given by|โ†‘ใ€‰ โŠ— |B๐›ผโ†‘ ใ€‰, again far away from resonance. From conservationof magnetization, the bath state now has total magnetization๐‘€ โˆ’ 1/2. Following the reset step of the central spin, this stateis reset to |โ†“ใ€‰ โŠ— |B๐›ผโ†‘ ใ€‰, which is no longer an eigenstate butrather a superposition of eigenstates. Crucially, these statesall have magnetization ๐‘€ โˆ’ 1: the total bath magnetizationhas been reduced. Dark states of the form |โ†“ใ€‰ โŠ— |D๐›ผโ†“ ใ€‰ are leftinvariant by these steps. After several cycles, dark state popula-tions build up and ultimately saturate the bath spin polarizationwell above its fully polarized value.

    The success of the protocol depends on the suppressionof diabatic excitations. However, transitions between brightstates within their own band are irrelevant for the purposes ofpolarization transfer, and thus we only require that transitionsbe suppressed between the bands. Specifically, we mimic aslow smooth ramp ๐œ†(๐‘ก) with ramp time ๐œ๐‘Ÿ ๏ฟฝ ๐œ0, where

    ๐œ0 = 2๐œ†0/ฮ”2min (18)

    sets the timescale for the onset of diabatic transitions betweeneigenstate bands (Appendix C). While such a protocol maystill be too slow in practical applications, here it serves only asa starting point which guarantees efficient transfer.

    III. POLARIZATION TRANSFER PROTOCOLS

    We detail how to speed up adiabatic ramps with the assis-tance of CD and LCD protocols in a single sweep. Such (L)CDprotocols can be exactly analyzed in the integrable limit. Wefurther compare our CD protocols to unassisted (UA) protocolswhich, unlike CD, attempt to polarize the bath without engi-neering additional controls. A full cooling protocol consistingof repeated sweeps will be analyzed in Section IV.

    We simulate sweeps ๐œ†(๐‘ก) across resonance by numericallysolving the time-dependent Schroฬˆdinger equation 3 in a specificpolarization sector and measure efficiency. The system isinitialized at ๐œ†๐‘– = โˆ’๐œ†0 ๏ฟฝ โˆ’ฮ”typ in a mixed state:

    ๐œŒ(๐‘ก = 0) = |โ†“ใ€‰ ใ€ˆโ†“| โŠ— ๐œŒ๐ต, (19)

    2 The reset and transfer steps have an effect on all polarization sectors simul-taneously.

    3 The specific ramp function used in this work is a smooth polynomial๐œ†(๐‘ก) = ๐œ†0 (12 (๐‘ก/๐œ๐‘Ÿ )5 โˆ’ 30 (๐‘ก/๐œ๐‘Ÿ )4 + 20 (๐‘ก/๐œ๐‘Ÿ )3 โˆ’ 1) , which monoton-ically increases from ๐œ†(0) = โˆ’๐œ†0 to ๐œ†(๐œ๐‘Ÿ ) = ๐œ†0 and has vanishing firstand second derivatives at the protocol boundaries. The minimal order ofa polynomial in ๐‘ก/๐œ๐‘Ÿ that satisfies these constraints is five. However, anyform ๐œ†(๐‘ก) with sufficiently smooth boundary conditions can be used [49].

  • 6

    where the bath is in an infinite-temperature state ๐œŒ๐ต. Thischoice of a spatially uncorrelated and unpolarized bath stateis motivated by experimental conditions. We also expect anycoherences in the initial bath state to be lost during the repeatedcycling of the qubit. This gives an initial probability ๐‘ƒ๐ตโ†“(๐‘ก =0) of starting in the top bright band and ๐‘ƒ๐ทโ†“(๐‘ก = 0) of startingin the dark band, with

    ๐‘ƒ๐ตโ†“,โ†‘(๐‘ก) =โˆ‘๏ธ๐›ผ

    Tr[๐œŒ(๐‘ก) Pโ†“,โ†‘ |B๐›ผใ€‰ใ€ˆB๐›ผ |Pโ†“,โ†‘], (20)

    ๐‘ƒ๐ทโ†“,โ†‘(๐‘ก) =โˆ‘๏ธ๐›ผ

    Tr[๐œŒ(๐‘ก) Pโ†“,โ†‘ |D๐›ผใ€‰ใ€ˆD๐›ผ |Pโ†“,โ†‘], (21)

    in which Pโ†“ โ‰ก |โ†“ใ€‰ ใ€ˆโ†“| โŠ— ๐ผ is the projection operator to thesubspace with down qubit polarization and similarly Pโ†‘ โ‰ก|โ†‘ใ€‰ ใ€ˆโ†‘| โŠ— ๐ผ. At the end of the ramp (๐œ† ๐‘“ = +๐œ†0), the state ๐œŒ(๐‘ก =๐œ๐‘Ÿ ) has a probability ๐‘ƒ๐ตโ†‘(๐‘ก = ๐œ๐‘Ÿ ) of being in the top brightband, ๐‘ƒ๐ทโ†“(๐‘ก = ๐œ๐‘Ÿ ) of being in the dark band, and ๐‘ƒ๐ตโ†“(๐œ๐‘Ÿ ) ofhaving transitioned to the bottom bright band. For protocolefficiency, we use two measures: (i) the transfer efficiency,

    ๐œ‚๐‘‡ โ‰ก ๐‘ƒ๐ตโ†‘(๐œ๐‘Ÿ )/๐‘ƒ๐ตโ†“(0), (22)

    which measures how effectively the qubit polarization in brightstates is flipped during a single sweep, and (ii) the kick effi-ciency,

    ๐œ‚๐พ โ‰ก 1 โˆ’ ๐‘ƒ๐ทโ†“(๐œ๐‘Ÿ )/๐‘ƒ๐ทโ†“(0), (23)

    which measures how effectively dark states are depopulated(or โ€˜kickedโ€™) into the bright manifold. Throughout this section,we continually refer to Fig. 3, which plots these efficienciesover a range of ramp times ๐œ๐‘Ÿ for numerically simulated UAand CD protocols. Note that we average over ๐‘๐‘  realizationsof disorder in ฮฉ๐ต, ๐‘— and ๐‘” ๐‘— , which we denote by an overline as๐œ‚๐‘‡ or ๐œ‚๐พ .

    A. Unassisted Driving (UA)

    We first discuss unassisted (UA) protocols, corresponding toa sweep of ๐œ† over a finite time. Adiabatic protocols correspondto infinite ramp times ๐œ๐‘Ÿ โ†’ โˆž, where all bright state polariza-tion is transferred across a single sweep: ๐œ‚๐‘‡ = 1 while ๐œ‚๐พ = 0.At finite ramp times diabatic effects become important andgenerally ๐œ‚๐‘‡ < 1 and ๐œ‚๐พ > 0. In the fast limit (๐œ๐‘Ÿ โ†’ 0) thesystem does not have time to respond to the drive, completelypreventing polarization transfer and dark state depletion suchthat ๐œ‚๐‘‡ , ๐œ‚๐พ โ†’ 0.

    In a system with a homogeneous bath field (๐›พ๐‘ง = 0), theoperator ๐‘†๐‘ง0 only couples bright state pairs

    4, and within each๐‘€ sector all excitations induced by a finite ramp speed ยค๐œ† > 0occur only between bright state pairs [34]. Each bright statepair can be treated as an independent two-level Landau-Zenerproblem following Eq. (8), for which the known transition

    4 Note that dark states at ๐›พ๐‘ง = 0 are eigenstates of ๐‘†๐‘ง0 , so they cannot coupleto bright states on changing the qubit z-field in time.

    FIG. 3. Efficiency vs. ramp time. Disorder-averaged transferefficiency (top) and kick efficiency (bottom) of UA, CD, and LCDprotocols in systems with ๐›พ๐‘ง = 0.00 (crosses) and ๐›พ๐‘ง = 0.05 (boxes).Dashed lines show the analytic prediction for the UA transfer effi-ciency (26), the analytic predictions for LCD transfer (41) and kickefficiencies (43) at large ramp velocities. Parameters: ๐‘๐‘  = 150,๐ฟ = 10, ๐‘€ = โˆ’1, ฮฉ๐ต = 10, ๐œ†0 = 5, ๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, and๐œ0 โ‰ˆ 1000.

    probability for a ramp ๐œ†(๐‘ก) across a resonant gap ฮ”๐›ผ is givenby [80, 81]

    ๐‘trans [ฮ”๐›ผ] = exp(โˆ’ ๐œ‹

    2ฮ”2๐›ผยค๐œ†

    ). (24)

    Averaging this transition probability over the gap distribution(14) returns an approximate transfer efficiency for a givenmagnetization sector

    ๐œ‚๐‘‡ = 1 โˆ’

    โˆซ โˆžฮ”min

    ๐‘trans [ฮ”] ๐‘›(ฮ”) ๐‘‘ฮ”โˆซ โˆžฮ”min

    ๐‘›(ฮ”) ๐‘‘ฮ”, (25)

    which can be evaluated to return

    ๐œ‚๐‘‡ = 1 โˆ’ยค๐œ†๐œ๐‘š

    1 + ยค๐œ†๐œ๐‘šexp

    (โˆ’๐‘š๐œ‹๐‘”

    2๐ฟ

    ยค๐œ†

    ), (26)

    in which ๐‘š = ๐‘€/๐ฟ is the magnetization density, ยค๐œ† โˆ ๐œ†0/๐œ๐‘Ÿ ,and

    ๐œ๐‘š =1

    ๐‘š๐œ‹๐‘”2๐ฟln

    (1 + 2๐‘š1 โˆ’ 2๐‘š

    ). (27)

  • 7

    The transfer efficiency in Eq. (26) is plotted in Fig. 3 as adashed black curve and shows excellent agreement with theUA calculations for ๐›พ๐‘ง = 0 and ๐›พ๐‘ง = 0.05.

    As shown in Fig. 3, the distinction between a system with ahomogeneous bath field (crosses) and an inhomogeneous bathfield (squares) has little impact on the UA transfer efficiency.The transfer efficiency in UA dynamics varies drastically with๐œ๐‘Ÿ . When ๐œ๐‘Ÿ is sufficiently large (๐œ๐‘Ÿ ๏ฟฝ ๐œ0; cf. Eq. (18)),transitions between eigenstate bands become suppressed andthe system becomes effectively adiabatic for the purposes ofpolarization transfer: the qubit flips in bright state bands, butnot in dark bands. Fig. 3 shows the tendency of simulated UAprotocols toward unit transfer efficiency.

    The difference between homogeneous and inhomogeneoussystems is important when considering the kick efficiency. Ina system with inhomogeneous bath fields (๐›พ๐‘ง > 0), ๐‘†๐‘ง0 couplesbright and dark eigenstates. As such, inhomogeneous fieldslead to a nonzero kick efficiency because diabatic transitionscan depopulate dark states, whereas homogeneous fields leadto a zero kick efficiency at all ramp rates.

    The convergence ๐œ‚๐‘‡ โ†’ 1 (shown) occurs much faster thanthe convergence ๐œ‚๐พ โ†’ 0+ (not shown). The former is deter-mined by the gap between bright state pairs, which remainfinite throughout the ramp at numerically accessible systemsizes, whereas the latter is determined by the dark-bright gaps,which tend to close away from resonance. This leads to dark-bright transitions at large yet finite ๐œ๐‘Ÿ . (ฮ”๐ธ)โˆ’2, where ฮ”๐ธ ison the order of the level spacing (Appendix C). Any attempt todrive the system faster (๐œ๐‘Ÿ . ๐œ0) leads to diabatic excitationsbetween eigenstates. When ๐›พ๐‘ง = 0, speeding up UA protocolsdecreases the transfer efficiency due to transitions betweenbright bands, but again does not deplete dark states. At finitedisorder strength ๐›พ๐‘ง = 0.05, UA protocols suffer a similar lossof transfer efficiency, but gain the ability to kick dark states,with a peak kick efficiency at intermediate speeds ๐œ๐‘Ÿ โˆผ ๐œ0.

    B. Exact Counterdiabatic Driving (CD)

    CD protocols suppress transitions between the eigenstatesof an instantaneous Hamiltonian by evolving the system withan assisted Hamiltonian that exactly cancels all diabatic ex-citations [49]. The inclusion of counterdiabatic terms in ahyperpolarization protocol can hence be used to increase thetransfer efficiency.

    Within each two-dimensional Landay-Zener subspace (8),the system remains in an instantaneous eigenstate of ๐ป๐›ผ (๐œ†(๐‘ก))at all times when evolved with a time-dependent Hamiltonian[49]

    ๐ปCD,๐›ผ (๐‘ก) = ๐ป๐›ผ (๐œ†(๐‘ก)) โˆ’ ยค๐œ†(๐‘ก)ฮ”๐›ผ

    ๐œ†(๐‘ก)2 + ฮ”2๐›ผ๐‘†๐‘ฆ๐›ผ, (28)

    where the auxiliary (counterdiabatic) term โˆ ๐‘†๐‘ฆ๐›ผ exactly can-cels diabatic transitions between the bright states for arbitraryramp speeds provided ยค๐œ† = 0 at the beginning and end of theramp.

    CD is realized for the full system if the system is evolvedwith the time-dependent CD Hamiltonian

    ๐ป๐ถ๐ท (๐‘ก) = ๐ป (๐œ†(๐‘ก)) + ยค๐œ†(๐‘ก) A๐œ† (๐œ†(๐‘ก)), (29)

    where the CD term A๐œ†, also known as the adiabatic gaugepotential, follows as

    A๐œ† (๐œ†) = โˆ’โˆ‘๏ธ๐›ผ

    ฮ”๐›ผ๐œ†2 + ฮ”2๐›ผ

    ๐‘†๐‘ฆ๐›ผ . (30)

    The summation index ๐›ผ runs over all bright pairs in all magne-tization sectors. The effect of the counterdiabatic term can beunderstood in the limit ยค๐œ† โ†’ โˆž, where the Hamiltonian reducesto ยค๐œ†A๐œ† and the evolution operator for a single sweep can bewritten as

    exp(๐‘–

    โˆซ โˆžโˆ’โˆž

    A๐œ† ๐‘‘๐œ†)=

    โˆ๐›ผ

    exp(โˆ’ ๐‘– ๐œ‹ ๐‘†๐‘ฆ๐›ผ

    ). (31)

    The gauge potential generates a rotation around the y-axis thatexchanges |B๐›ผ+ ใ€‰ โ†” |B๐›ผโˆ’ ใ€‰ when ๐œ† is swept across resonance,exactly as happens in the adiabatic protocol.

    Alternatively, the gauge potential can be written in closedform as (see Appendix F)

    A๐œ† = โˆ’๐‘–

    4(๐ป โˆ’ฮฉ๐ต๐‘€)โˆ’2 [๐ป, ๐‘†๐‘ง0] . (32)

    The first (inverse) term in the product is to be interpreted in thesense of a pseudo-inverse, and the second (commutator) termin the product is given by:

    [๐ป, ๐‘†๐‘ง0] = ๐‘–โˆ‘๏ธ๐‘—

    ๐‘” ๐‘— (๐‘†๐‘ฅ0 ๐‘†๐‘ฆ

    ๐‘—โˆ’ ๐‘†๐‘ฆ0 ๐‘†

    ๐‘ฅ๐‘— ). (33)

    The gauge potential is a complex many-body operator, dif-ficult to compute in theory and even harder to implement inpractice [49]. Only in certain special cases, for example when๐œ•๐œ†๐ป is an integrable perturbation of an integrable model ๐ป,is this operator sufficiently local [49, 82, 83]. Fortunately,๐œ•๐œ†๐ป = ๐‘†

    ๐‘ง0 is an integrable perturbation of ๐ป in the ๐›พ๐‘ง = 0 limit

    of our present model, and the pair structure of the bright statecould be used to immediately write down the adiabatic gaugepotential. A similar two-level structure for the gauge potentialalso arises in integrable free-fermionic systems [82].

    In a system with an inhomogeneous bath field (๐›พ๐‘ง > 0), wecan no longer express the adiabatic gauge potential explicitly.Nevertheless, as CD mimics an adiabatic protocol, the transferefficiency will be maximal.

    Fig. 3 showcases the effect of exact CD in a transfer proto-col across resonance for systems with ๐›พ๐‘ง = 0 and ๐›พ๐‘ง = 0.05(crosses and squares respectively). In both cases, the completesuppression of bright state transitions yields a maximally ef-ficient transfer protocol ๐œ‚๐‘‡ = 1, systematically improving onthe UA protocol, while the complete suppression of dark statetransitions results in zero kick efficiency ๐œ‚๐พ = 0.

    C. Local Counterdiabatic Driving (LCD)

    In practice, it is hard to realize exact CD, and we mustresort to approximation schemes. In this section, we follow the

  • 8

    method devised in Ref. [55] to develop a local approximationALCD to A๐œ†. We refer to assisted driving (see Eq. (29)) withALCD as local counterdiabatic driving (LCD).

    As proposed in Ref. [55] and detailed in Appendix G, aformal expansion for the adiabatic gauge potential can be foundin terms of nested commutators:

    A๐œ† = ๐‘–๐‘žโˆ‘๏ธ๐‘—=1๐›ผ ๐‘— [๐ป, [๐ป, . . . [๐ป๏ธธ ๏ธท๏ธท ๏ธธ

    2 ๐‘—โˆ’1

    , ๐œ•๐œ†๐ป]]] . (34)

    For ๐‘ž โ†’ โˆž Eq. (34) reproduces the exact gauge potential. Alocal approximation for A๐œ† is obtained by truncating the com-mutator expansion of Eq. (34) to a desired order ๐‘ž, and using avariational minimization scheme [54] to set the coefficients ๐›ผ ๐‘—for ๐‘— = 1, . . . , ๐‘ž.

    We focus on the leading order term because it (i) is simpleenough to be implemented by Floquet driving on the qubit field(see Section VI) and (ii) is already remarkably effective forpolarization transfer. One can always refine the approximationto CD by adding higher-order commutators in Eq. (34); therapid convergence of higher-order LCD to CD is shown inAppendix G.

    To leading order (๐‘ž = 1), we obtain:

    ALCD (๐œ†) = ๐‘– ๐›ผ1 (๐œ†) [๐ป, ๐‘†๐‘ง0], ๐›ผ1 (๐œ†) = โˆ’1

    ๐œ†2 + ฮ”2typ, (35)

    This scheme leads to a coefficient ๐›ผ1 (๐œ†) depending on a singleenergy scale which coincides exactly with the typical gap ฮ”typ,in contrast with exact CD, where the prefactor depends eitherexplicitly (28) or implicitly (32) on all bright state gaps ฮ”๐›ผ. Insum,

    ๐ปLCD (๐‘ก) = ๐ป (๐œ†(๐‘ก)) + ๐‘– ยค๐œ†(๐‘ก) ๐›ผ1 (๐œ†(๐‘ก)) [๐ป, ๐‘†๐‘ง0]

    = ๐ป (๐œ†(๐‘ก)) +ยค๐œ†(๐‘ก)

    ๐œ†(๐‘ก)2 + ฮ”2typ

    โˆ‘๏ธ๐‘—

    ๐‘” ๐‘— (๐‘†๐‘ฅ0 ๐‘†๐‘ฆ

    ๐‘—โˆ’ ๐‘†๐‘ฆ0 ๐‘†

    ๐‘ฅ๐‘— ).

    (36)

    Fig. 3 shows the resulting LCD curves as unmarked solidred curves (๐›พ๐‘ง = 0.0), and circle-marked red curves (๐›พ๐‘ง =0.05). LCD is approximate, ๐œ‚๐‘‡ < 1 (see top panel of Fig. 3).Nevertheless, LCDโ€™s transfer efficiency is high (๐œ‚๐‘‡ & 0.75)over the whole range of ramp times ๐œ๐‘Ÿ , even as ๐œ๐‘Ÿ โ†’ 0 whereUA becomes completely transfer inefficient.

    A finer comparison between ALCD and A๐œ† can be made for๐›พ๐‘ง = 0 by expressing the gauge potential in the Landau-Zenerpicture (8),

    ALCD (๐œ†) = โˆ’โˆ‘๏ธ๐›ผ

    (ฮ”๐›ผฮ”typ

    )ฮ”typ

    ๐œ†2 + ฮ”2typ๐‘†๐‘ฆ๐›ผ . (37)

    Rather than targeting individual gaps as in Eq. (30), the LCDprotocols effectively target a single typical energy splittingscale to suppress diabatic transitions between the bright bands.In contrast with Eq. (30), the Lorentzian prefactor of ๐‘†๐‘ฆ๐›ผ hasa fixed width ฮ”typ, which does not vary with bright state gap,and a modulated amplitude ฮ”๐›ผ/ฮ”typ.

    This discrepancy introduces polarization transfer errors inLCD at intermediate and fast ramps (๐œ๐‘Ÿ . ๐œ0). Comparing withEq. (31), in the limit ยค๐œ† โ†’ โˆž the LCD protocol again generatesa rotation within bright state pairs:

    exp(๐‘–

    โˆซ โˆžโˆ’โˆž

    A๐ฟ๐ถ๐ท ๐‘‘๐œ†)=

    โˆ๐›ผ

    exp(โˆ’ ๐‘– ๐œ‹ ฮ”๐›ผ

    ฮ”typ๐‘†๐›ผ๐‘ฆ

    ). (38)

    LCD strongly suppresses transitions between those bright pairswith a gap ฮ”๐›ผ โ‰ˆ ฮ”typ, but otherwise yield only partial suppres-sion.

    From Eq. (38) we can define a mismatch error between CDand LCD for each bright pair with gap ฮ”๐›ผ as

    E[ฮ”๐›ผ] = cos2(๐œ‹

    2ฮ”๐›ผฮ”typ

    ). (39)

    Averaging over the gap distribution (14), the transfer efficiencyat large ramp rates is

    ๐œ‚๐‘‡ = 1 โˆ’

    โˆซ โˆžฮ”min

    E[ฮ”] ๐‘›(ฮ”) ๐‘‘ฮ”โˆซ โˆžฮ”min

    ๐‘›(ฮ”) ๐‘‘ฮ”, (40)

    The saddle-point approximation returns

    ๐œ‚๐‘‡ =

    โˆซ โˆž1 ๐‘‘๐‘ก ๐‘ก sin

    2(๐œ‹2โˆš

    2๐‘š๐‘ก) (

    1โˆ’2๐‘š1+2๐‘š

    ) ๐‘ก2โˆซ โˆž1 ๐‘‘๐‘ก ๐‘ก

    (1โˆ’2๐‘š1+2๐‘š

    ) ๐‘ก2 . (41)This expression agrees with the LCD transfer efficiency inFig. 3 (dashed red line) and will be discussed in more detail inthe following section.

    Fig. 3 (bottom panel) also shows the LCD kick efficiencyover several ๐œ๐‘Ÿ orders. In the ๐›พ๐‘ง = 0 limit, LCD has no effecton dark states, just like UA and CD, again leading to a zerokick efficiency (crosses in bottom panel of Fig. 3). When๐›พ๐‘ง > 0, LCD protocols do not prevent dark-bright transitionsand exhibit non-zero kick efficiency. Since the bright-dark gapis smaller than the typical bright band gap ฮ”typ, especially farfrom resonance where the bright-dark gap tends to close, LCDallows dark-bright transitions as in UA driving.

    The difference in gap scales gives LCD both the advantagesof CD for efficient transfer, and the advantages of diabatic UAfor depopulating dark states. For slow ramps ๐œ๐‘Ÿ > ๐œ0, LCD andUA have similar transfer efficiencies as the diabatic transitionprobabilities are small. In faster ramps (๐œ๐‘Ÿ . ๐œ0), bright bandtransitions are suppressed by LCD but not UA. Meanwhile,LCD saturates to a maximum kick efficiency for ๐œ๐‘Ÿ < ๐œ0, incontrast with UA protocols which peak around ๐œ๐‘Ÿ โˆผ ๐œ0 and thenlose kick efficiency as ๐œ๐‘Ÿ โ†’ 0. The distinction between LCDand UA protocols in this fast limit will be further quantified inEq. (43), following Appendix E, where it is argued that in thelimit ๐œ๐‘Ÿ โ†’ 0 the kick efficiency is proportional to the transferefficiency.

    Finally, note that this work focuses on the weak xx-disorderlimit ๐›พ๐‘ฅ๐‘ฅ < ๐‘” where there is a finite gap between bright bandsat numerically accessible system sizes. However, a finite brightpair gap is not necessary to design efficient LCD protocols that

  • 9

    need only target a typical gap between the bright bands. InAppendix D we show that LCD maintains high transfer andkick efficiencies in the presence of strong xx-disorder evenin the presence of small gap as long as the bulk of the brightspectrum still has a gap โˆผ ฮ”typ.

    D. Protocol Efficiency and Polarization Sector

    We compare the efficiencies of CD, LCD, and UA protocolsfor a single sweep across different polarization sectors ๐‘€.Since all protocols systematically reduce polarization, it iscrucial to understand how the transfer and kick efficienciesdepend on the polarization sector. The results are summarizedin Fig. 4 for a fast ramp (๐œ๐‘Ÿ = 0.05 ๐œ0) in a system with aninhomogeneous bath field (๐›พ๐‘ง = 0.05) for multiple systemsizes ๐ฟ.

    FIG. 4. Efficiency vs. polarization sector. The vertical axes showtransfer efficiency (top) and kick efficiency (bottom) for fast LCD(colored) and UA (grey) sweeps across resonance in a system withan inhomogeneous bath field. The horizontal axis shows the density๐‘/๐ฟ = ๐‘€/๐ฟ + 1/2 of spin flips above the fully polarized state in๐‘€ < 0 sectors. We plot theoretical predictions to ๐œ‚๐‘‡ based on thethermodynamic limit calculations at zero disorder (cf. Eqs. (26) and(41)), and to ๐œ‚๐พ based on ๐œ‚๐‘‡ (cf. Eq. (43)). For reference, wealso plot corresponding efficiencies for CD protocols. Parameters:๐‘๐‘  = 150, ฮฉ๐ต = 10, ๐œ†0 = 5, ๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, and ๐œ๐‘Ÿ = 500/๐ฟ.

    As expected, CD always produces unit transfer efficiencyand zero kick efficiency. Moreover, LCD outperforms UA

    by both efficiency measures in every polarization sector. Thetop panel shows the transfer efficiency ๐œ‚๐‘‡ , plotted against๐‘/๐ฟ = ๐‘€/๐ฟ โˆ’ 1/2.For both LCD and UA protocols, the trans-fer efficiency decreases with ๐‘/๐ฟ because the minimal gapand the number of bright state pairs ๐‘›๐ต =

    ( ๐ฟโˆ’1๐‘โˆ’1

    )increases with

    ๐‘/๐ฟ, in turn increasing the likelihood of diabatic transitionsbetween bright pairs. For LCD, the transfer efficiency is max-imal (๐œ‚๐‘‡ = 1) in the sector with ๐‘ = 1 because there is onlyone bright state pair with gap ฮ”typ = ฮ”LCD to target.

    The bottom panel shows the kick efficiency ๐œ‚๐พ , plottedagainst ๐‘/๐ฟ. For both LCD and UA, the kick efficiency in-creases with polarization. This increase can be understoodby comparing the number ๐‘›๐ท of dark states to the number ofbright pairs ๐‘›๐ต within each sector. In the sector with ๐‘ = 1,there are (๐ฟ โˆ’ 2) dark states compared to a single pair ofbright states, which severely limits the pool of bright states thatdark states can transition to. As we probe increasingly larger๐‘ , ๐‘›๐ต eventually surpasses ๐‘›๐ท such that ๐‘›๐ต/๐‘›๐ท โ†’ O(๐ฟ)as ๐‘/๐ฟ โ†’ 0.5. The number of available bright states thatdark states can transition to increases and thus enhances kickefficiency.

    Fig. 4 also shows the analytic predictions (black curves) forthe transfer efficiency from Eq. (26) and (41), consistent withthe collapse of the curves. For LCD in fast ramps (41), the onlydependence of ๐œ‚๐‘‡ is on ๐‘š = ๐‘€/๐ฟ, consistent with the collapseof ๐œ‚๐‘‡ curves at different system sizes as a function of spinflip density (๐‘/๐ฟ โˆผ ๐‘€/๐ฟ + 1/2). The transfer efficiency inEq. (26) depends on both ๐‘š and ๐‘”2๐ฟ/ ยค๐œ†. To achieve a collapseof curves at different system sizes, one must also scale ยค๐œ† โˆผ ๐ฟto eliminate the residual ๐ฟ dependence, yielding a transfer effi-ciency which depends only on ๐‘/๐ฟ. In practice, the collapsecan be achieved by scaling up ๐œ†0 โˆผ ๐ฟ at fixed ramp-time ๐œ๐‘Ÿ orscaling down ๐œ๐‘Ÿ โˆผ 1/๐ฟ at fixed ramp range; in our simulations,we have implemented the latter. Both predictions show excel-lent agreement with simulation results in most magnetizationsectors, except near ๐‘ โˆผ O(1) where finite size effects aresignificant. Such finite-size effects also lead to the deviation ofthe numerically observed bright pair gap distribution in Fig. 2from the analytical one at large negative values of ๐‘€ .

    Remarkably, there is a simple approximate relation between๐œ‚๐‘‡ and ๐œ‚๐พ for LCD and UA protocols in the presence ofz-disorder at moderate-to-fast ramps ๐œ . ๐œ0. Along withEqs. (41) and (26), this relation provides an analytical predic-tion for the kick efficiency, such that the protocol efficiencycan be fully characterized analytically. Assume that the proba-bility weight that is not successfully transferred to states withup qubit polarization ergodically mixes between the availabledark and bright states with spin down. Then,

    ๐‘ƒ๐ทโ†“(๐œ๐‘Ÿ ) โ‰ˆ ๐‘ƒ๐ทโ†“(0)๐‘›๐ต (1 โˆ’ ๐œ‚๐‘‡ ) + ๐‘›๐ท

    ๐‘›๐ต + ๐‘›๐ท, (42)

    which implies a kick efficiency

    ๐œ‚๐พ = 1 โˆ’๐‘ƒ๐ทโ†“(๐œ๐‘Ÿ )๐‘ƒ๐ทโ†“(0)

    โ‰ˆ ๐‘›๐ต๐‘›๐ต + ๐‘›๐ท

    ๐œ‚๐‘‡ =๐‘

    (๐ฟ โˆ’ ๐‘) ๐œ‚๐‘‡ . (43)

    The black curves in Fig. 4 show ๐œ‚๐พ computed using Eqs. (43),(41) and (26). These analytic curves are in good agreement

  • 10

    with the numerical data for both LCD and UA. Note that thederivation of Eq. (43) assumes equal mixing between dark andbright states: closer to integrable points, this assumption breaksdown, and an analytic relation between the transfer and kickefficiency is no longer possible.

    The resulting protocols hence depend on the interplay ofdifferent effects across magnetization sectors: increasing ๐‘š,the total number of bright states capable of transfering polar-ization increases, the average transfer efficiency decreases, andthe kick efficiency increases.

    IV. HYPERPOLARIZING THE BATH

    FIG. 5. Polarizing the spin bath. Expectation value of the bathpolarization per spin ใ€ˆใ€ˆ๐‘†๐‘ง

    ๐ตใ€‰ใ€‰ along a sequence of ๐‘c = 100 back and

    forth cycles of the detuning ๐œ† across resonance. Top and bottom panelsshow a typical realization for systems with zero and finite z-disorder,respectively. The system is initialized in an infinite temperature state.The forward-backward transfer flow between resets is illustrated bythe gold arrows. Grey dotted lines denote the polarization of the fullypolarized state. Parameters: ๐ฟ = 4, ฮฉ๐ต = 10, ๐œ†0 = 5, ๐‘” = 0.1,๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0 (top), ๐›พ๐‘ง = 0.05 (bottom), and ๐œ๐‘Ÿ /๐œ0 โ‰ˆ 0.05.

    We now turn to the performance of the various protocolsover multiple reset-transfer cycles. In particular, we show howthe ability of LCD to kick dark states enables complete bathspin polarization.

    Figs. 5 and 6 illustrate the progressive polarization of thebath over multiple (๐‘๐‘ = 100) reset-transfer cycles in UA, CD,

    FIG. 6. Spin bath polarization vs. cycle. Average bath polarizationper spin vs. cycle number with the same setup as in Fig. 5. Parameters:๐‘๐‘  = 1, ๐ฟ = 4, ฮฉ๐ต = 10, ๐œ†0 = 5, ๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0 (top),๐›พ๐‘ง = 0.05 (bottom), and ๐œ๐‘Ÿ /๐œ0 โ‰ˆ 0.05.

    and LCD protocols. We focus on fast sweeps ๐œ๐‘Ÿ/๐œ0 โ‰ˆ 0.05,where the effects of LCD and UA are significantly differen-tiated. For this simple demonstration, we consider a qubitcoupled to 3 bath spins; however, the observed qualitativebehavior generalizes to larger baths (see Section V). In bothfigures, we measure the expectation value of the average bathspin polarization per spin:

    ใ€ˆใ€ˆ๐‘†๐‘งBใ€‰ใ€‰ =1

    ๐ฟ โˆ’ 1๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    ใ€ˆ๐‘†๐‘ง๐‘—ใ€‰, (44)

    where ใ€ˆ๐‘†๐‘ง๐‘—ใ€‰ โ‰ก Tr[๐œŒ(๐‘ก)๐‘†๐‘ง

    ๐‘—] is the expectation value of ๐‘†๐‘ง

    ๐‘—in the

    density matrix ๐œŒ(๐‘ก) of the system at time ๐‘ก.In Fig. 5, the bath polarization per spin is shown as a func-

    tion of the detuning ๐œ†(๐‘ก) across resonance. After each transfersweep, the qubit polarization is reset and the direction of theramp reversed; the resulting forward-backward motion is de-picted by the gold arrows. Fig. 6 shows the corresponding bathpolarization per spin after every cycle.

    In a typical realization of a system with a homogeneousbath field (๐›พ๐‘ง = 0), CD protocols at first quickly reduce thebath polarization due to their maximal transfer efficiency, butsoon slow down and saturate as dark states become populated.The saturation point lies well above the fully polarized state(see grey dotted line ใ€ˆใ€ˆ๐‘†๐‘งBใ€‰ใ€‰ = โˆ’0.5). In contrast, UA protocolsare relatively inefficient and much slower to reach saturation,requiring many more sweeps. LCD protocols perform onlyslightly worse than CD and much better than UA; they eventu-ally also saturate above the fully polarized state.

    In a typical realization of a system with an inhomogeneousbath field (๐›พ๐‘ง = 0.05), CD protocols behave the same as in thehomogeneous limit, quickly polarizing the bath to a saturationpoint. LCD protocols no longer saturate and can polarize thebath close to the fully polarized state due to their non-zero kickefficiency. Since the hyperpolarization scheme progressivelypopulates smaller ๐‘€ sectors, and the kick efficiency decreaseswith decreasing ๐‘€ (see Fig. 4), the polarization rate per cycledecreases as we ใ€ˆใ€ˆ๐‘†๐‘ง

    ๐ตใ€‰ใ€‰ โ†’ โˆ’1/2. UA protocols, like LCD, are

  • 11

    able to fully polarize the bath, but their smaller kick efficiencyrequires many more sweeps.

    V. SCALING TO LARGE BATHS

    In this section, we explore how the number of cycles neededto hyperpolarize the bath scales with system size ๐ฟ. So farwe focused on relatively small system sizes ๐ฟ . 10 to designand test our protocols with accessible exact dynamic simu-lations. To circumvent the resource cost of simulating exactdynamics with larger system sizes, we introduce a scalablemaster equation of the hyperpolarization process in terms ofprobability flow equations which should be accurate at largetransfer speeds.

    The state of the system after ๐‘ transfer-reset polarizationcycles is given by:

    ๐‘ƒ(๐‘) =[ยฎ๐‘ƒ๐ตยฎ๐‘ƒ๐ท

    ]=

    ๐‘ƒ๐ต,โ†“[0]๐‘ƒ๐ต,โ†“[1]

    ...

    ๐‘ƒ๐ต,โ†“[๐ฟ]๐‘ƒ๐ท,โ†“[0]๐‘ƒ๐ท,โ†“[1]

    ...

    ๐‘ƒ๐ท,โ†“[๐ฟ]

    (45)

    where ๐‘ƒ๐ต,โ†“[๐‘] is the probability of finding the system in abright state with down qubit polarization in the sector with ๐‘spin flips above the fully polarized state, and ๐‘ƒ๐ท,โ†“[๐‘] is theprobability of finding the system in a dark state with downqubit polarization in the same sector. We do not track the prob-abilities of bright and dark states with up qubit polarization, asthey are always converted to states with down qubit polariza-tion after reset. Moreover, as there exist no dark states withdown qubit polarization for ๐‘€ โ‰ฅ 0, ๐‘ƒ๐ท,โ†“[๐‘] = 0 for ๐‘ โ‰ฅ ๐ฟ/2.Finally, we assume that the bath is fully characterized by theprobabilities in Eq. (45) and ignore any correlations in thedensity matrix between individual dark and bright states sincethe bath generally decoheres between different polarizationcycles. This assumption is justified a posteriori by compar-ing the efficiencies predicted by the master equation to exactsimulations.

    The dynamics of the system is obtained by applying a trans-fer matrix ๐‘‡ :

    ๐‘ƒ(๐‘ + 1) = ๐‘‡ ๐‘ƒ(๐‘), (46)

    where the transfer matrix can be schematically written as

    ๐‘‡ =

    [๐‘‡๐ต๐ต ๐‘‡๐ต๐ท

    ๐‘‡๐ท๐ต ๐‘‡๐ท๐ท

    ]. (47)

    The transfer efficiency ๐œ‚๐‘‡ sets the probability that the qubitpolarization is flipped in bright states during a sweep acrossresonance. On the other hand, the kick efficiency ๐œ‚๐พ sets the

    FIG. 7. Effective model. Schematic representing the action of thetransfer matrix through the efficiency functions ๐œ‚ and reset rates ๐‘Ÿin bright (red) and dark (black) manifolds in two neighboring po-larization sectors ๐‘ = ๐‘– + 1 and ๐‘ = ๐‘– during a single polarization(transfer+reset) cycle.

    probability that dark states are kicked into bright states. Weassume that dark states with down qubit polarization are onlykicked into bright states with down qubit polarization. Whenthe qubit polarization is reset after each sweep, bright stateswith qubit state |โ†‘ใ€‰ transition to either bright states (with |โ†“ใ€‰)or dark states (with |โ†“ใ€‰) in a lower magnetization sector, withrelative probability ๐‘Ÿ๐ต and ๐‘Ÿ๐ท , respectively. Therefore, thenon-zero matrix elements of the transfer matrix are given by:

    ๐‘‡๐ต๐ต [๐‘–, ๐‘–] = 1 โˆ’ ๐œ‚๐‘‡ [๐‘–] (48)๐‘‡๐ต๐ต [๐‘– โˆ’ 1, ๐‘–] = ๐‘Ÿ๐ต [๐‘–] ๐œ‚๐‘‡ [๐‘–] (49)๐‘‡๐ท๐ต [๐‘– โˆ’ 1, ๐‘–] = ๐‘Ÿ๐ท [๐‘–] ๐œ‚๐‘‡ [๐‘–] (50)๐‘‡๐ท๐ท [๐‘–, ๐‘–] = 1 โˆ’ ๐œ‚๐พ [๐‘–] (51)๐‘‡๐ต๐ท [๐‘–, ๐‘–] = ๐œ‚๐พ [๐‘–] (52)

    for every sector index ๐‘– = 0, . . . , ๐ฟ. Fig. 7 illustrates the trans-fer and reset rates for a single cycle. Note ๐‘Ÿ๐ต [๐‘–] + ๐‘Ÿ๐ท [๐‘–] = 1,so only one reset rate needs to be specified.

    As shown in Section III, the different protocols CD, UA, andLCD have different efficiency functions ๐œ‚๐‘‡ [๐‘–] and ๐œ‚๐พ [๐‘–]. Here,we consider moderate-to-fast ramp speeds (๐œ๐‘Ÿ . ๐œ0) whereLCD and UA have distinct effects. Since we are interestedin obtaining the scaling of the full protocol, we linearize theprevious expressions and model the kick efficiency for LCDand UA as

    ๐œ‚๐พ [๐‘–] = ๐œ‚0๐‘–

    ๐ฟ; (๐‘– โ‰ค ๐ฟ/2), (53)

    and model the corresponding transfer efficiency ๐œ‚๐‘‡ [๐‘–] (๐‘– โ‰ค๐ฟ/2) using Eq. (43).

    The reset rates ๐‘Ÿ๐ท [๐‘–] and ๐‘Ÿ๐ต [๐‘–] depend on the probabilitydistribution of the state within each bright/dark band and detailsof the structure of eigenstates. Furthermore, these rates candrastically change from one disorder realization to another andare hence difficult to predict. To get a reasonable estimate forour master equation, we take ๐‘Ÿ๐ท [๐‘–+1] and ๐‘Ÿ๐ต [๐‘–+1] proportionalto to the number of accessible dark and bright states in the ๐‘–๐‘กโ„Ž

    sector, respectively. For ๐‘€ โ‰ฅ 0, all the weight is transferred tobright states,

    ๐‘Ÿ๐ต [๐‘– + 1] = 1, ๐‘€ โ‰ฅ 0, (54)

  • 12

    since dark states have qubit spin up and are not accessibleduring reset. For ๐‘€ < 0, dark states have qubit spin down andbecome accessible, such that

    ๐‘Ÿ๐ต [๐‘– + 1] =๐‘›๐ต [๐‘–]

    ๐‘›๐ต [๐‘–] + ๐‘›๐ท [๐‘–], ๐‘€ < 0, (55)

    with ๐‘›๐ท and ๐‘›๐ต given by Eqs. (9) and (16), and we refer to thisreset rate as well-mixed. Although we cannot generally satisfythe equiprobable condition within every bright band, we expectto approximately have well-mixed reset rates on average inlarge disordered systems.

    We test the master equation in Fig. 8, which compares UA,CD, and LCD dynamics with Eq. (46) to the correspondingexact dynamics for different system sizes (๐ฟ = 4, 8). The figureplots the average bath polarization per spin over many cyclesfor a single disorder realization. Our master equation agreeswell with the results from exact dynamics for all protocols,justifying the assumptions in Eq. (45). The exact dynamicsis computed at small ramp times ๐œ๐‘Ÿ โ‰ˆ 0.05 ๐œ0. To properlycapture protocol efficiencies at this ramp speed, we set ๐œ‚0 โ‰ˆ 1.0for LCD and ๐œ‚0 โ‰ˆ 0.4 for UA in accordance with the results inFig. 4.

    FIG. 8. Spin bath polarization vs. cycle. Average bath spinpolarization after every transfer-reset cycle over many cycles. Theleft and right panels show simulation results for systems of size ๐ฟ = 4and ๐ฟ = 8, respectively. Colored markers indicate numerical resultsusing our scalable master equation. Solid colored lines correspond toexact dynamics simulations. Parameters: ๐‘๐‘  = 1, ฮฉ๐ต = 10, ๐œ†0 = 5,๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0.05, and ๐œ๐‘Ÿ = 500/๐ฟ.

    The master equation allows acces to much larger systemsizes compared to exact diagonalization. Fig. 9 shows thenumber of cycles ๐‘๐‘ required to reach 99% of the polarizationof the fully polarized state against system size ๐ฟ. The opaqueand faint curves denote master equations capturing ๐œ๐‘Ÿ = 0.01 ๐œ0and ๐œ๐‘Ÿ = 0.05 ๐œ0, respectively.

    We find that the number of polarization cycles needed tofully polarize the bath in both LCD and UA protocols scaleslinearly with system size ๐ฟ. The main difference between LCDand UA is in the prefactor, depending on protocol duration, andwhich can be orders of magnitude larger in the UA protocolcompared to the LCD protocol at sufficiently fast ramps. Inslower ramps ๐œ๐‘Ÿ > ๐œ0 (not shown), LCD and UA have similar

    prefactor, but the prefactor for UA however increases as ๐œ๐‘Ÿis decreased. Our results are consistent with the expectationthat as ๐œ๐‘Ÿ โ†’ 0, UA takes progressively more cycles to fullypolarize the bath. Thus, moderate-to-fast LCD is not only time-efficient but also optimizes the number of cycles required toreach the fully polarized state.

    We conclude this section with a couple of remarks. (i)The master equation is applicable at sufficiently fast ramptimes ๐œ๐‘Ÿ < ๐œ0 โˆผ ๐œ†0 ฮ”โˆ’2min, where ฮ”min โˆผ

    โˆš๐ฟ ๐‘” in the lowest

    polarization sectors. To ensure this condition holds as ๐ฟ โ†’ โˆž,we scale ๐œ†0 โˆผ ๐ฟ. Otherwise at fixed ๐œ†0 and sufficiently large๐ฟ โˆผ ๐œ†0/(๐œ๐‘Ÿ ๐‘”2), the master equation would need to be refinedto properly account for more complicated speed dependenciesin the transfer and kick efficiencies. (ii) Similarly, our masterequation is based on efficiency measurements at sufficientlylarge z-disorder, where ๐›พ๐‘ง & ฮ”2min/๐œ†0. In the lowest energysectors, this requires ๐›พ๐‘ง & ๐ฟ ๐‘”2/๐œ†0. Again, the ๐ฟ dependencecan be cancelled by scaling ๐œ†0 โˆผ ๐ฟ.

    VI. FLOQUET ENGINEERING (FE) OF LCD

    The physical implementation of the LCD protocol requiresrealizing a non-trivial operator [๐ป, ๐‘†๐‘ง0]. We show how it ispossible to obtain this LCD Hamiltonian as an effective high-frequency Hamiltonian through Floquet engineering.

    Floquet engineering focuses on the design and physical ef-fects of periodic drives [84]. A periodically driven systemexhibits dynamics which can be described stroboscopicallyusing an effective slow/static Floquet Hamiltonian ๐ป๐น . Fre-quently, a control is periodically modulated at a frequencyscale ๐œ” larger than any other dynamical frequency in the sys-tem, and ๐ป๐น can be (Magnus) expanded in powers of ๐œ”โˆ’1 [84].In addition to capturing high-frequency physics, the Magnus

    FIG. 9. Number of cycles to 99% polarization vs. system size.Master equation simulation results are shown for UA (dashed black)and LCD (solid red) protocol. Opaque curves show results for ๐œ‚0 =1.0 set for LCD and ๐œ‚0 = 0.1 set for UA, which model ramps with๐œ๐‘Ÿ = 0.01 ๐œ0. At this ramp speed, we find ๐‘๐‘ โ‰ˆ 4๐ฟ for LCD and๐‘๐‘ โ‰ˆ 40๐ฟ for UA. Faint curves show results for ๐œ‚0 = 1.0 set for LCDand ๐œ‚0 = 0.4 set for UA, which model ramps with ๐œ๐‘Ÿ = 0.05 ๐œ0. Atthis ramp speed, we find ๐‘๐‘ โ‰ˆ 4๐ฟ for LCD and ๐‘๐‘ โ‰ˆ 10๐ฟ for UA.

  • 13

    expansion has a commutator structure closely related to thestructure of the gauge potential in Eq. (34), and can be used torealize local counterdiabatic driving at every order [55].

    A. Two-level system

    In order to give some intuition for the many-body Floquetprotocol, we illustrate the general ideas on the two-level sys-tem of Eqs. (8) and (28). Temporarily dropping the brightstate label and making the time-dependence implicit, the CDHamiltonian to be realized can be written as

    ๐ปCD = ๐œ†๐‘†๐‘ง + ฮ”๐‘†๐‘ฅ + ยค๐œ†๐›ผ1ฮ”๐‘†๐‘ฆ . (56)

    In an experimental set-up only ๐œ† is an accessible control pa-rameter, whereas ฮ” is constant and the ๐‘†๐‘ฆ term is absent (asit corresponds to a complex many-body operator acting onthe bright pair states). In order to realize ๐ปCD as an effectiveHamiltonian, we consider a LZ Hamiltonian and add high-frequency oscillations modulated by a slowly-varying ampli-tude. Specifically, we consider a time-dependent Hamiltonianof the form

    ๐ปFE (๐‘ก) = ๐›พ(๐‘ก)๐‘†๐‘ง + ฮ”๐‘†๐‘ฅ+

    [๐›ฝ(๐‘ก)๐œ” sin(๐œ”๐‘ก) + ยค๐›ฝ(๐‘ก) (1 โˆ’ cos(๐œ”๐‘ก))

    ]๐‘†๐‘ง , (57)

    with ๐›ฝ(๐‘ก) and ๐›พ(๐‘ก) slowly-varying functions to be determined.The choice of this time-dependence is motivated by the re-

    sulting effective Hamiltonian: in the limit of a large driving fre-quency ๐œ”, the stroboscopic dynamics for this time-dependentHamiltonian is generated by the Floquet Hamiltonian (derivedin Appendix H)

    ๐ปF = ๐›พ๐‘†๐‘ง + ๐ฝ0 (๐›ฝ)ฮ”[cos(๐›ฝ) ๐‘†๐‘ฅ โˆ’ sin(๐›ฝ) ๐‘†๐‘ฆ

    ], (58)

    where the slow time-dependence has been made implicit and๐ฝ0 is a Bessel function of the first kind.

    The effective Hamiltonian is of the form (56), containing a๐‘†๐‘ฆ term not present in the instantaneous Hamiltonian. How-ever, in the CD Hamiltonian the prefactor of ๐‘†๐‘ฅ is constantand the prefactor of ๐‘†๐‘ฆ is time dependent. Since the (slow)time dependence of these terms in the Floquet Hamiltonian isdetermined by the same factor ๐›ฝ(๐‘ก), it is not possible to directlyrealize the CD Hamiltonian in this way. Rather, we can realizea Hamiltonian proportional to the CD Hamiltonian.

    Demanding ๐ป๐น = ๐บ (๐‘ก)๐ป๐ถ๐ท , the prefactor for ๐‘†๐‘ฅ immedi-ately returns the time-dependent prefactor of the full Hamilto-nian as

    ๐บ (๐‘ก) = ๐ฝ0 (๐›ฝ(๐‘ก)) cos(๐›ฝ(๐‘ก)). (59)

    Time evolution follows the time-dependent Schroฬˆdinger equa-tion

    ๐‘–๐œ•๐‘ก |๐œ“(๐‘ก)ใ€‰ = ๐บ (๐‘ก)๐ปCD |๐œ“(๐‘ก)ใ€‰ . (60)

    Defining a โ€˜rescaled timeโ€™ ๐‘ (๐‘ก) such that ๐œ•๐‘  = ๐บ (๐‘ก)๐œ•๐‘ก , Eq. (60)can be used to realize counterdiabatic control in the rescaled

    time provided ๐‘–๐œ•๐‘  |๐œ“(๐‘ก (๐‘ ))ใ€‰ = ๐ปCD (๐‘ (๐‘ก)) |๐œ“(๐‘ (๐‘ก))ใ€‰. The coun-terdiabatic term is obtained by setting

    tan(๐›ฝ(๐‘ก)) = โˆ’๐›ผ1 (๐‘ (๐‘ก)) ยค๐œ†(๐‘ (๐‘ก)), (61)

    determining ๐›ฝ(๐‘ก) as function of ๐›ผ1 (๐‘ก), leaving

    ๐›พ(๐‘ก) = ๐บ (๐‘ก)๐œ†(๐‘ (๐‘ก)), (62)

    to finally return ๐ป๐น = ๐บ (๐‘ก)๐ปCD (๐‘ (๐‘ก)). Note that the experi-mental time necessarily runs in the positive direction, requiring๐บ (๐‘ก) > 0 and ๐›ฝ โˆˆ [โˆ’๐œ‹/2, ๐œ‹/2].

    B. FE protocol

    The ideas in Section VI A can be immediately extended tothe many-body Hamiltonian and LCD of Eq. (36). Given atarget LCD ramp with ๐œ†(๐‘ก) = ฮฉ๐‘„ (๐‘ก) โˆ’ฮฉ๐ต, we drive the systemwith the Floquet engineered (FE) Hamiltonian:

    ๐ปFE = ๐ป (ฮ›(๐‘ก)), (63)

    with a modified field detuning ฮ›(๐‘ก) = ฮฉ๐‘„ (๐‘ก) โˆ’ฮฉ๐ต given by

    ฮ›(๐‘ก) = ๐ฝ0 (๐›ฝ(๐‘ก)) cos(๐›ฝ(๐‘ก)) ๐œ†(๐‘ (๐‘ก))+ ๐›ฝ(๐‘ก) ๐œ” sin(๐œ” ๐‘ก) + ยค๐›ฝ(๐‘ก) (1 โˆ’ cos(๐œ” ๐‘ก)). (64)

    Following Eq. (61), we set

    ๐›ฝ(๐‘ก) โ‰ก arctan(โˆ’ ๐‘‘๐œ†(๐‘ (๐‘ก))

    ๐‘‘๐‘ ๐›ผ1 (๐‘ (๐‘ก))

    ), (65)

    and the rescaled time ๐‘  = ๐‘ (๐‘ก) satisfying ๐‘‘๐‘  = ๐บ (๐‘ก)๐‘‘๐‘ก is definedas

    ๐‘  =โˆซ ๐‘ก

    0๐ฝ0 (๐›ฝ(๐‘ก โ€ฒ)) cos(๐›ฝ(๐‘ก โ€ฒ)) ๐‘‘๐‘ก โ€ฒ > 0. (66)

    This FE Hamiltonian is designed precisely so that the leadingorder approximation to its Floquet Hamiltonian ๐ป๐น in thehigh-frequency limit is the LCD Hamiltonian in the rescaledtime:

    ๐ปF = ๐ป (๐œ†(๐‘ ))+๐‘–๐‘‘๐œ†(๐‘ )๐‘‘๐‘ 

    ๐›ผ1 (๐‘ ) [๐ป (๐œ†(๐‘ )), ๐œ•๐œ†๐ป]+O(

    1๐œ”

    ). (67)

    More specifically, the effective Floquet Hamiltonian is foundas (see Appendix H)

    ๏ฟฝฬƒ๏ฟฝF =๐บ (๐‘ก)[๐œ†(๐‘ (๐‘ก)) ๐‘†๐‘ง0 +

    โˆ‘๏ธ๐‘—

    ๐›ฟฮฉโ€ฒ๐‘— ๐‘†๐‘ง๐‘—

    +โˆ‘๏ธ๐‘—

    ๐‘” ๐‘— (๐‘†๐‘ฅ0 ๐‘†๐‘ฅ๐‘— + ๐‘†

    ๐‘ฆ

    0 ๐‘†๐‘ฆ

    ๐‘—) โˆ’ tan(๐›ฝ(๐‘ก)) ๐‘– [๐ป, ๐‘†๐‘ง0]

    ], (68)

    where ๐›ฟฮฉโ€ฒ๐‘—โ‰ก (ฮฉ๐ต, ๐‘— โˆ’ฮฉ๐ต)/๐บ (๐‘ก) is the renormalized z-

    disorder.The FE protocol is stroboscopically equivalent to LCD with

    ALCD in Eq. (35). Moreover, in a smooth ramp ๐œ† with ยค๐œ†๐‘– =ยค๐œ† ๐‘“ = 0, ๐ป๐น๐ธ and ๐ป๐น yield the exact same initial and final

  • 14

    states, which guarantees that FE and LCD produce the samepolarization transfer during our hyperpolarization scheme.

    We remark that ๐ป๐น equals ๐ปLCD only in the absence of z-disorder (๐›พ๐‘ง = 0). At finite z-disorder (๐›พ๐‘ง > 0), the two differdue to the renormalization ๐›ฟฮฉโ€ฒ

    ๐‘—of the bath fields. Away from

    this point, the renormalization tends to enhance z-disordersince ๐บ (๐‘ก) โˆˆ [0, 1]. No significant quantitative differences inperformance were found between FE (in lab time) and LCD(in rescaled time) with renormalized disorder, as shown next.

    FIG. 10. Floquet engineered ramp. Top panel shows the FE rampฮ›(๐‘ก) (solid colored curves) as a function of time ๐‘ก for ramp times๐œ๐‘Ÿ = 0.025, 0.5, 1.0. The target ramp ๐œ†(๐‘ก) (dashed black curve) isshown for reference. The bottom panel shows the effect of FE rampson the mean qubit polarization ใ€ˆ๐‘†๐‘ง0 (๐‘ก)ใ€‰ over the course of the ramp at๐œ๐‘Ÿ = 0.25, 0.5. The corresponding LCD curves are shown to coincidewith the FE curves. Curves for UA and CD at ๐œ๐‘Ÿ = 0.05 ๐œ0 areshown for reference. Parameters: ๐ฟ = 8, ฮฉ๐ต = 10, ฮ›0 = ๐œ†0 = 5,๐‘” = 0.1, ๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0.05, ๐œ0 โ‰ˆ 1000, and ๐œ” = 100 (Note: Fordisplay, we have graphically reduced ๐œ” by a factor 10 to decreasecurve density).

    The upper panel of Fig. 10 showcases the FE ramp ฮ›(๐‘ก)in Eq. (64) for various ramp times ๐œ๐‘Ÿ . The vertical axis is re-scaled by the magnitude of the initial/final detunings ฮ›0 โ‰ก ๐œ†0,which are designed to coincide with the target ramp at the rampendpoints. The target ramp ๐œ†(๐‘ก) is shown for reference (dashedblack curve). Near the adiabatic breakdown time ๐œ๐‘Ÿ/๐œ0 = 1,the FE ramp ฮ›(๐‘ก) has a base profile (averaging out the oscil-lations) similar to ๐œ†(๐‘ก), with small oscillation amplitudes thatget slightly more pronounced in the middle of the ramp aroundresonance. For progressively faster ramps ๐œ๐‘Ÿ/๐œ0 = 0.05, 0.025,

    the FE ramp ฮ›(๐‘ก) shows more pronounced deviations from๐œ†(๐‘ก). First observe that the base profile of FE changes, keepingthe system near resonance for a progressively larger amountof time. Moreover, the amplitude of the high-frequency oscil-lations around resonance progressively increases due to ๐›ฝ(๐‘ก)in ฮ›(๐‘ก). Physically, these properties ensure the qubit and bathinteract strongly and long enough to effect polarization transferin accordance with LCD.

    The lower panel of Fig. 10 serves two purposes: (i) to showthe effect of FE on the mean qubit z-polarization ใ€ˆ๐‘†๐‘ง0ใ€‰ overthe course of a sweep, and (ii) to highlight the equivalenceof FE and LCD protocols. The qubit is initialized with spindown ใ€ˆ๐‘†๐‘ง0ใ€‰ = โˆ’0.5 in a mixed state. Over the course of theramp ฮ›(๐‘ก), FE (solid colored curves) transfers a large fractionof the qubit polarization to the bath in perfect agreement withLCD (dashed white lines). For reference we also show an UAprotocol at ๐œ๐‘Ÿ/๐œ0 = 0.05 (dashed black curve); as expected it ismuch less efficient compared to FE/LCD and CD at this rampspeed.

    In sum, we can systematically realize LCD with FE, wherethe LCD protocol can be implemented indirectly in experi-ments by driving the local qubit field ฮฉ๐‘„ (๐‘ก) periodically athigh-frequencies ๐œ” ๏ฟฝ ฮฉ๐‘„,ฮฉ๐ต, ๐œโˆ’1๐‘Ÿ . Importantly, the FE pro-tocol requires no controls which are not already present in ๐ปin Eq. (2), similar in spirit to Ref. [59]. It can be achieved bysetting a fixed global field ฮฉ๐ต and dynamically varying ฮฉ๐‘„ (๐‘ก),without modifying the qubit-bath interactions. This result dif-fers from other schemes which require controlling interactionsto realize LCD with Floquet engineering [55, 57, 58].

    C. Quantum Speed Limit

    The distinction between the lab time ๐‘ก and the rescaled time๐‘  gives rise to a quantum speed limit. Namely, there exists acritical ramp time ๐œ๐‘Ÿ = ๐œ๐‘†๐ฟ > 0 in the lab frame for whichthe protocol duration ๐œ๐‘† in rescaled time becomes zero and๐›ฝ โ†’ ๐œ‹/2. Given sufficiently large driving frequencies, it isalways possible to realize LCD using FE if the LCD ramp timeis larger than this critical ramp time. However, at shorter ramptimes, the proposed protocol would lead to negative protocoldurations in stretched times, and the FE protocol can no longerrealize LCD. The speed limit can derived (see Appendix I) byinverting Eq. (66):

    ๐œ๐‘†๐ฟ = lim๐‘ โ†’0

    โˆซ ๐‘ 0

    [๐บ (๐‘ก (๐‘ โ€ฒ)]โˆ’1๐‘‘๐‘ โ€ฒ โˆผ ฮ”โˆ’1typ. (69)

    The timescale ๐œ๐‘†๐ฟ is set by the typical bright pair gap ฮ”typ,which is on the order of the time needed to transfer polarizationfrom the qubit to the bath while sitting at resonance. In fact,the profile ๐บ (๐‘ก)๐œ†(๐‘ (๐‘ก)) of the FE protocol in Eq. (63) reducesto a sudden quench protocol to resonance as ๐œ โ†’ ๐œ๐‘†๐ฟ .

    Eqs. (64) and (69) provide an additional connection be-tween sudden and adiabatic polarization protocols: Floquet-engineering implements the adiabatic polarization protocol ina transformed frame, which resembles a sudden protocol inthe lab frame when the speed limit is approached. For rampsfaster than the speed limit ๐œ๐‘Ÿ/๐œ๐‘†๐ฟ < 1, FE can be extended

  • 15

    FIG. 11. Power vs. protocol time. Disorder-averaged transfer poweragainst the scaled ramp time ๐œ๐‘Ÿ /๐œ0 for UA, CD, and FE protocols. Thevertical axis is scaled by (1/๐œ๐‘†๐ฟ) (grey dash-dotted line). The verticalblue solid line marks the disorder-averaged speed limit timescale ๐œ๐‘†๐ฟ .Parameters: ๐‘๐‘  = 50, ๐ฟ = 8, ฮฉ๐ต = 10, ฮ›0 = ๐œ†0 = 5, ๐‘” = 0.1,๐›พ๐‘ฅ๐‘ฅ = 0.05, ๐›พ๐‘ง = 0.05, ๐œ0 โ‰ˆ 1000, and ๐œ” = 100.

    by taking ฮ›(๐‘ก) = ๐œ‹2 ๐œ” sin(๐œ” ๐‘ก); then FE effectively oscillatesaround resonance for a shorter time than ๐œ๐‘†๐ฟ and is no longeras effective as LCD.

    This speed limit is quantified in Fig. 11, showing the transferpower vs. ramp time for several transfer protocols acrossresonance. The transfer power is defined as

    ฮ”๐‘†๐‘ง0๐œ๐‘Ÿ

    =ใ€ˆ๐‘†๐‘ง0 (๐œ† ๐‘“ )ใ€‰ โˆ’ ใ€ˆ๐‘†

    ๐‘ง0 (๐œ†๐‘–)ใ€‰

    ๐œ๐‘Ÿ, (70)

    measuring the rate at which polarization is extracted from thequbit per unit ramp time. In the absence of tunable qubit-bathinteractions, the maximum possible polarization transfer is aunit of polarization on the timescale ๐œ๐‘†๐ฟ , shown in the plot asa grey dash-dotted line. We find that UA protocols operate farbelow this rate over the whole range of ramp times. On theother hand, FE protocols significantly enhance power, peakingin the vicinity of the speed limit. At protocol durations ๐œ > ๐œ๐‘†๐ฟFE nears the efficiency of CD protocols which, as expected,transfers polarization more effectively than UA and FE perunit time. At protocol durations ๐œ < ๐œ๐‘†๐ฟ , the FE power liessignificantly below the CD power.

    The presence of this speed limit suggests a broader physicallimitation: Any ramped protocol which does not directly tunesystem-bath couplings or add extra controls cannot transferpolarization at a faster rate than a sudden resonant exchange.

    VII. CONCLUSION

    In this work, we apply the tools of shortcuts to adiabaticityto a class of hyperpolarization protocols. In a single cycle ofeach such protocol, the qubit is reset along the โˆ’๐‘ง direction byan external pulse, after which the ๐‘ง-field of the qubit is sweptacross a resonance region. Polarization is transferred from thequbit to the spin bath during the sweep.

    We introduce local counterdiabatic driving (LCD) proto-cols that mimic an adiabatic protocol. The LCD protocols

    simultaneously tackle two problems: (i) the small sweep ratesnecessary for the adiabatic transfer of polarization to the bath,and (ii) the limits on hyperpolarization imposed by dark states.The LCD protocols tackle (i) by efficiently suppressing dia-batic transitions between bright bands. They tackle (ii) bydepleting dark states in the presence of inhomogeneous bathfields (i.e., when the system is non-integrable). In this way,LCD protocols outperform both unassisted protocols and exactcounterdiabatic protocols since the former does not suppresstransitions between bright bands and the latter suppresses tran-sitions from dark to bright bands.

    Using exact numerics and a master equation, we show thatthe LCD protocols outperform the unassisted ones by variousmetrics (efficiency, power, and number of cycles). Addition-ally, the LCD can be experimentally implemented through ahigh-frequency Floquet drive on the qubit. These engineeredprotocols have a natural quantum speed limit; once the sweeprate exceeds this limit, the LCD protocols cannot be realizedthrough Floquet drives.

    The LCD may be used to speed up hyper-polarization inseveral experimental systems with dipolarly interacting spins.Indeed, Eq. (2) models the (rotating-frame) Hamiltonian of ashallow nitrogen-vacancy (NV) defect coupled to surface elec-tronic spins in high-purity diamond [79, 85, 86], as well as the(rotating-frame) Hamiltonian of a NV defect coupled to bulkC-13 nuclei [7, 12] or the nuclei of external molecules in solu-tion [17]. A promising avenue for future work is to compare theperformance of LCD protocols to sudden protocols that satisfythe Hartmann-Hahn condition [87] in these systems. Hyper-polarization of powdered diamond using NV-centers [12, 14]is also an important goal for magnetic resonance imaging. Itwould be interesting to develop LCD protocols that account forthe random orientation of the NV center axes (the ๐‘ง-directionof the central spin in Eq. (2)) in these systems. An additionaldirection for future work is to use optimal control theory todesign possibly more efficient protocols and test the validityof the speed limit. However, such a numerical optimizationproblem in the many-body setting is expected to be highlycomplex. In contrast, the LCD approach, while not guaranteedto be optimal, is readily extended to more complex systemswith an arbitrary number of spins and arbitrary interactions.

    Theoretically, our work raises questions about the precise in-terplay between integrability and hyperpolarizability in centralspin models. The model in Eq. (2) with ๐›พ๐‘ง = 0 is (i) integrable,and (ii) has exact dark states for any choice of ๐‘” ๐‘— [34]. How-ever, the closely related XXX model with ๐›พ๐‘ง = 0 and isotropicqubit-bath interactions

    โˆ‘๐‘— ๐‘” ๐‘— ยฎ๐‘†0 ยท ยฎ๐‘† ๐‘— is integrable without ex-

    hibiting dark states [88, 89]. The XXX model describes thehyperfine interactions of the electronic spin of a quantum dotwith surrounding nuclei [90, 91]. Previous work [31, 92] sug-gests that the spin bath can be efficiently polarized in the XXXmodel despite its integrability. A natural direction for futurework is to quantify the general role of integrability in the po-larization process. Another possible direction is to examinethe role of interactions in the spin bath. The accompanyingdiffusive spin transport is expected to aid in the polarization ofdistant bath spins with negligible ๐‘” ๐‘— .

  • 16

    ACKNOWLEDGEMENTS.

    The authors thank E. Boyers, M. Pandey, C.R. Laumann,and A. Sushkov for insightful discussions. The authors ac-knowledge support from the Sloan Foundation through a SloanResearch Fellowship (A.C.), from the Belgian American Edu-cational Foundation (BAEF) through the Francqui FoundationFellowship (P.W.C.), and from the BU CMT Visitor Program(P.W.C.). Numerics were performed on the BU Shared Com-puting Cluster with the support of the BU Research Comput-ing Services. This work was supported by EPSRC Grant No.EP/P034616/1 (P.W.C.), NSF DMR-1813499 (T.V. and A.P.),NSF DMR-1752759 and AFOSR FA9550-20-1-0235 (T.V. andA.C.), and AFOSR FA9550-16- 1-0334 (A.P.).

    Appendix A: Derivation of central spin Hamiltonian

    Consider a driven qubit-bath spin system in a magnetic field๐ต along the z-direction described by the Hamiltonian:

    ๐ป0 = ๐ป๐‘„ + ๐ป๐ต + ๐ป๐ท + ๐ป๐‘„๐ต + ๐ป๐ต๐ต, (A1)

    Here ๐ป๐‘„ = ๐›พ๐‘„ ๐ต ๐‘†๐‘ง0 is the Zeeman energy of the central qubitwith gyromagnetic ratio ๐›พ๐‘„, and ๐ป๐ต = ๐›พ๐ต ๐ต

    โˆ‘๐ฟโˆ’1๐‘—=1 ๐‘†

    ๐‘ง๐‘—

    is theZeeman energy of the bath spins with gyromagnetic ratio ๐›พ๐ต.The driving term is given by

    ๐ป๐ท = 2ฮฉ๐‘„ cos(๐œ”๐‘„๐‘ก) ๐‘†๐‘ฅ0 + 2 cos(๐œ”๐ต๐‘ก)๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    ฮฉ๐ต, ๐‘— ๐‘†๐‘ฅ๐‘— (A2)

    where ฮฉ๐‘„ is a Rabi amplitude on the central spin, ฮฉ๐ต, ๐‘— =ฮฉ๐ต + ๐›ฟฮฉ ๐‘— is a generally inhomogeneous Rabi amplitude on thebath, and ๐œ”๐‘„, ๐œ”๐ต are respectively the corresponding drivingfrequencies. The qubit-bath coupling is given by a dipoleinteraction term:

    ๐ป๐‘„๐ต =โˆ‘๏ธ๐‘–

    ๐›พ๐‘„๐›พ๐ต

    ๐‘Ÿ3๐‘–

    [ยฎ๐‘†0 ยท ยฎ๐‘†๐‘– โˆ’ 3 ( ยฎ๐‘†0 ยท ๐‘Ÿ๐‘–) ( ยฎ๐‘†๐‘– ยท ๐‘Ÿ๐‘–)

    ](A3)

    where ยฎ๐‘Ÿ๐‘– is the vector between the central qubit and the ๐‘–thbath spin. The interaction term ๐ป๐ต๐ต between bath spin pairsis generally also dipolar. In this work, we assume bath-bathinteractions are small compared to the qubit-bath couplings,which is realized in experiments with sufficiently low bathspin density or with bath spins that satisfy ๐›พ๐ต ๏ฟฝ ๐›พ๐‘„. TheHamiltonian ๐ป0 in Eq. (A1) describes single qubit systems,such as NV centers in diamond or quantum dots, interactingwith an ensemble of spins (e.g. spins on the surface of di-amond) and driven by continuous irradiation fields (such asradio waves) [5, 16, 17, 19, 23, 24, 38].

    In a doubly rotated frame defined by the unitary transforma-tion

    ๐‘ˆ = exp[โˆ’ ๐‘–

    (๐œ”๐‘„ ๐‘†

    ๐‘ง0 + ๐œ”๐ต

    ๐ฟโˆ’1โˆ‘๏ธ๐‘–=1

    ๐‘†๐‘ง๐‘–

    )๐‘ก

    ], (A4)

    ๐ป0 can be simplified by matching the driving frequencies tothe Zeeman energies (๐œ”๐‘„ = ๐›พ๐‘„๐ต, ๐œ”๐ต = ๐›พ๐ต๐ต), and applying

    a rotating wave approximation to eliminate rapidly rotatingnon-secular terms which average to zero on the timescale ofthe dynamics [24]. Relabeling our axes (๐‘ฅ, ๐‘ง) โ†’ (๐‘ง,โˆ’๐‘ฅ), thedominant time-averaged motion is described by

    ๐ปrot (๐‘ก) = ฮฉ๐‘„๐‘†๐‘ง0 +๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    ฮฉ๐ต, ๐‘—๐‘†๐‘ง0 +

    ๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    2 ๐‘” ๐‘— ๐‘†๐‘ฅ0 ๐‘†๐‘ฅ๐‘— (A5)

    where

    ๐‘” ๐‘— โ‰ก๐›พ๐‘„๐›พ๐ต

    2 ๐‘Ÿ3๐‘–

    [1 โˆ’ 3 cos2 (๐œƒ๐‘–)], (A6)

    and ๐œƒ๐‘– is the angle between ยฎ๐ต and ยฎ๐‘Ÿ๐‘– in the frame of ๐ป0.We note our rotating wave approximation requires |๐‘” ๐‘—/๐ต | ๏ฟฝ๐›พ๐‘„ , ๐›พ๐ต , |๐›พ๐‘„ ยฑ ๐›พ๐ต | , which is readily satisfied in NV centers orquantum dot experiments [16, 18].

    The interaction term

    ๐‘†๐‘ฅ0 ๐‘†๐‘ฅ๐‘— =

    14(๐‘†+0๐‘†

    โˆ’๐‘— + ๐‘†โˆ’0 ๐‘†

    +๐‘— + ๐‘†+0๐‘†

    +๐‘— + ๐‘†โˆ’0 ๐‘†

    โˆ’๐‘—

    )(A7)

    describes zero quantum (flip-flop) transitions in its first twoterms, and double quantum (flip-flip/flop-flop) transitions in-teractions in its last two terms. Zero quantum transitions dom-inate when ๐‘” ๐‘— ๏ฟฝ ฮฉ๐‘„ + ฮฉ๐ต, ๐‘— [24], yielding the Hamiltonianpresented in the main text:

    ๐ป (๐‘ก) = ฮฉ๐‘„ ๐‘†๐‘ง0 +๐ฟโˆ’1โˆ‘๏ธ๐‘—=1

    ฮฉ๐ต, ๐‘— ๐‘†๐‘ง๐‘—+ 1

    2

    ๐ฟโˆ’1โˆ‘๏ธ๐‘—=1๐‘” ๐‘—

    (๐‘†+0๐‘†

    โˆ’๐‘— + ๐‘†โˆ’0 ๐‘†

    +๐‘—

    ). (A8)

    Appendix B: Distribution of energy gaps in the homogeneouslimit

    In this section, we compute the approximate distribution ofbright pair gaps ฮ”๐›ผ (Eq.(14)) in a system with a homogeneousbath field (๐›พ๐‘ง = 0).

    In the homogeneous limit (๐›พ๐‘ฅ๐‘ฅ = 0), the central spin modelreduces to a two-body Hamiltonian

    ๐ป = ๐œ†๐‘†๐‘ง0 +๐‘”

    2(๐‘†+0๐‘†

    โˆ’ + ๐‘†โˆ’0 ๐‘†+) , (B1)

    with ๐‘†ยฑ =โˆ‘๐ฟโˆ’1๐‘—=0 ๐‘†

    ยฑ๐‘—. The spectrum of this Hamiltonian can be

    obtained in the collective bath spin basis, as the Hamiltonianonly couples the states: |โ†‘ใ€‰ โŠ— |๐‘ , ๐‘šใ€‰ , |โ†“ใ€‰ โŠ— |๐‘ , ๐‘š + 1ใ€‰. Here ๐‘  isthe total spin quantum number of the bath, and ๐‘š is the totalz-projection of the bath state, leading to ๐‘€ = ๐‘š + 1/2. We takeโˆ’๐‘  < ๐‘š < ๐‘ , since the states |โ†‘ใ€‰ โŠ— |๐‘ , ๐‘ ใ€‰ and |โ†“ใ€‰ โŠ— |๐‘ ,โˆ’๐‘ ใ€‰ aredark eigenstates of the Hamiltonian.

    The energy in each two-dimensional subspace is fully deter-mined by the quantum numbers ๐‘  and ๐‘š, leading to energiesยฑฮ”๐‘ ,๐‘š/2 at resonance given by

    ฮ”๐‘ ,๐‘š = ๐‘”โˆš๏ธ(๐‘  โˆ’ ๐‘š) (๐‘  + ๐‘š + 1). (B2)

    The number of gaps equal to ฮ”๐‘ ,๐‘š is fully determined by thenumber of ways the (๐ฟโˆ’1) spin-1/2 bath spins can be coupled

  • 17

    to a collective spin ๐‘  with spin projection ๐‘š. Furthermore,since ๐‘š < ๐‘  and ๐‘š is fixed by specifying ๐‘€, this also leadsto a minimal gap within each polarization sector ๐‘€ , given byฮ”๐‘š = ๐‘”

    โˆš2๐‘€ + 1, obtained by setting ๐‘  = ๐‘š + 1 = ๐‘€ + 1/2.

    Increasing ๐‘  resulting in an increasing ฮ”๐‘ ,๐‘š, whereas smallervalues of ๐‘  are not allowed within this polarization sector.

    Given (๐ฟ โˆ’ 1) bath spins, the number of spin-๐‘  representa-tions is given by Catalanโ€™s triangle as

    ๐‘๐‘  (๐ฟ) = ๐ถ ((๐ฟ โˆ’ 1)/2 + ๐‘ , (๐ฟ โˆ’ 1)/2 โˆ’ ๐‘ ) (B3)

    =(๐ฟ โˆ’ 1)!(2๐‘  + 1)

    (๐ฟ/2 โˆ’ 1/2 โˆ’ ๐‘ )!(๐ฟ/2 + 1/2 + ๐‘ )! , (B4)

    which can be approximated for large ๐ฟ as

    ๐‘๐‘  (๐ฟ) โ‰ˆ๐‘’๐ฟ ๐‘“๐‘  (1/2โˆ’๐‘ /๐ฟ)

    โˆš2๐œ‹

    2๐‘  + 1๐ฟ/2 + ๐‘  + 1

    โˆš๏ธ„๐ฟ

    (๐ฟ/2 โˆ’ ๐‘ ) (๐ฟ/2 + ๐‘ ) ,

    (B5)with ๐‘“๐‘  (๐‘) = โˆ’๐‘ ln(๐‘) โˆ’ (1 โˆ’ ๐‘) ln(1 โˆ’ ๐‘). The total magneti-zation ๐‘€ fixes ๐‘š such that the energy gap only depends on ๐‘ ,and we can introduce a gap density as

    ๐‘›(ฮ”) = ๐‘๐‘  (ฮ”) (๐ฟ)๐‘‘๐‘ 

    ๐‘‘ฮ”, (B6)

    with ๐‘ (ฮ”) =โˆš๏ธ(ฮ”/๐‘”)2 + ๐‘€2 โˆ’ 1/2 and

    ๐‘‘๐‘ 

    ๐‘‘ฮ”=

    ฮ”/๐‘”2โˆš๏ธ(ฮ”/๐‘”)2 + ๐‘€2

    . (B7)

    As mentioned before, every fixed magnetization sector has aminimal gap, which we write as ฮ”๐‘š โ‰ก ๐‘”

    โˆš2๐‘€ + 1, such that

    ๐‘›(ฮ” < ฮ”๐‘š) = 0. Due to the presence of the exponentialterm, in the limit of large ๐ฟ all integrals over ๐‘›(ฮ”)๐‘‘ฮ” willbe dominated by the boundary terms where ฮ” โ‰ˆ ฮ”๐‘š. Ap-proximating the exponential factor at ๐‘ /๐ฟ = |๐‘€ |/๐ฟ โ‰ก ๏ฟฝฬƒ๏ฟฝ forฮ”๐‘š < ฮ” ๏ฟฝ ๐‘”๐‘€ , we find

    ๐‘›(ฮ”) โ‰ˆ ๐พ (๏ฟฝฬƒ๏ฟฝ) ๐‘’๐ฟ ๐‘“๐‘  (1/2โˆ’๏ฟฝฬƒ๏ฟฝ)โˆš

    2๐œ‹๐ฟฮ”

    ฮ”2๐‘š

    (1 โˆ’ 2๏ฟฝฬƒ๏ฟฝ1 + 2๏ฟฝฬƒ๏ฟฝ

    ) ฮ”2ฮ”2๐‘š, (B8)

    with

    ๐พ (๏ฟฝฬƒ๏ฟฝ) = 4๏ฟฝฬƒ๏ฟฝ1/2 + ๏ฟฝฬƒ๏ฟฝ

    โˆš๏ธ„1

    1/4 โˆ’ ๏ฟฝฬƒ๏ฟฝ2. (B9)

    Appendix C: Diabatic transitions in the Landau-Zener problem

    The Landau-Zener (LZ) problem, described in Eq. (8) ofthe main text, consists of a two-level system with gap ฮ”๐ฟ๐‘ =โˆš๐œ†2 + ฮ”2, where ๐œ† is a control field and ฮ” is the minimum

    gap [80, 81].When the control field is varied at a speed ยค๐œ† โˆผ ๐œ†0/๐œ๐‘Ÿ , we

    can estimate the speed scale below which the system remainsadiabatic. Adiabaticity occurs when the rate of change of thegap ฮ” is smaller than the dynamical timescale over the whole

    range of the control field ๐œ†. In particular, this condition holdsif it is satisfied near resonance (|๐œ† | โˆผ ฮ”) where the gap issmallest. ( ยคฮ”๐ฟ๐‘

    ฮ”๐ฟ๐‘๏ฟฝ ฮ”๐ฟ๐‘

    ) ๏ฟฝ๏ฟฝ๏ฟฝ๏ฟฝ๐œ†=ฮ”

    =โ‡’ ยค๐œ† ๏ฟฝ ฮ”2. (C1)

    Therefore we satisfy the adiabatic condition when the ramptimescale ๐œ๐‘Ÿ ๏ฟฝ ๐œ†0/ฮ”2. In faster ramps moreover, the scale๐œ0 โˆผ ๐œ†0/ฮ”2 sets the scale for the onset of diabatic transitions.

    As discussed in the main text for our central spin model, theLZ problem directly captures the interactions between brightbands. Nevertheless, the LZ problem can also help understandtransitions between dark and bright bands, as we discuss next.

    In system with ๐›พ๐‘ง = 0, the gap ฮ”๐ท๐ต between a dark statewith energy ๐ธ๐ท and a neighboring bright state with energy ๐ธ๐ตis given by

    ฮ”๐ท๐ต = |๐ธ๐ต โˆ’ ๐ธ๐ท | =12

    โˆš๏ธƒ๐œ†2 + ฮ”2min ยฑ

    12๐œ†, (C2)

    since ๐ธ๐ท = ยฑ๐œ†/2 and ๐ธ๐ต = ยฑ 12โˆš๏ธƒ๐œ†2 + ฮ”2min and where

    ฮ”min โ‰ก min๐›ผฮ”๐›ผ is the minimum bright-bright gap at reso-nance. At resonance, the gap is ฮ”๐ท๐ต = ฮ”min/2, comparableto the minimum bright-bright gap. In contrast with bright-bright gaps, dark-bright gaps are smallest furthest away fromresonance ๐œ† = ยฑ๐œ†0:

    minฮ”๐ท๐ต โ‰ˆ14

    (ฮ”min๐œ†0

    )2๐œ†0. (C3)

    In systems with ๐›พ๐‘ง = 0, the presence of a small bright-dark gapdoes not imply fast ramps yield diabatic transitions becausethe driving operator ๐‘†๐‘ง0 does not couple bright a